acceptorless alcohol dehydrogenation: a mechanistic

19
REVIEW Acceptorless Alcohol Dehydrogenation: A Mechanistic Perspective Pragati Pandey 1 Indranil Dutta 1 Jitendra K. Bera 1 Received: 1 August 2016 / Accepted: 18 August 2016 Ó The National Academy of Sciences, India 2016 Abstract Alcohols are unreactive and require strong inorganic oxidants to convert to synthetically useful car- bonyl compounds. Acceptorless dehydrogenation of alco- hol is a green and atom-economic alternative, which provides aldehyde (or ketone) without the use of sacrificial acceptor molecules and the side product is molecular hydrogen. This review provides a brief overview of the initial work followed by recent advances in the field of acceptorless alcohol dehydrogenation. Catalysts that employ metal–ligand cooperation for alcohol activation and dehydrogenation are covered in details. Different mechanisms are examined and clear advantages associated with a bifunctional pathway are outlined. Mechanistic understanding at the molecular level helps to develop new generation dehydrogenation catalysts. Recent works from our group on this area along with literature reports are discussed. Keywords Acceptorless alcohol dehydrogenation Bifunctional catalysis Metal–ligand cooperation Dehydrogenation mechanism Bifunctional double hydrogen transfer Abbreviations A ˚ Angstrom AAD Acceptorless alcohol dehydrogenation AD Alcohol dehydrogenation ADHC Acceptorless dehydrogenative coupling Bn Benzyl bMepi 1,3-Bis(6-methyl-2-pyridylimino)isoindolate KO t Bu Potassium tert-butoxide BDHT Bifunctional double hydrogen transfer bpyO a,a 0 -Bipyridonate Cp Cyclopentadienyl Cp* Pentamethylcyclopentadienyl Cy Cyclohexyl DABCO 1,4-Diazabicyclo[2.2.2]octane DBU 1,8-Diazabicyclo[5.4.0]undec-7-ene dppf 1,1 0 -Bis(diphenylphosphino)ferrocene DFT Density functional theory eu Energy unit Et Ethyl GC–MS Gas chromatography–mass spectrometry HMB Hexamethyl benzene IMes 1,3-Bis(2,4,6-trimethylphenyl)imidazol-2- ylidene I i Pr 1,3-Diisopropyl imidazol-2-ylidene i Pr Isopropyl MLC Metal ligand cooperation Me Methyl n Bu n-Butyl NHC N-Heterocyclic carbine NTs N-Tosyl Nu Nucleophile OAc Acetate OTf Trifluoromethanesulphonate Ph Phenyl phenO 2,9-Dihydroxy-1,10-phenanthroline t Bu Tertiarybutyl TOF Turn-over-frequency TEMPO 2,2,6,6-Tetramethylpiperidine-1-oxyl Tp 0 Tris(3,5-dimethylpyrazolyl)borate & Jitendra K. Bera [email protected] 1 Department of Chemistry, Center for Environmental Sciences and Engineering, Indian Institute of Technology Kanpur, Kanpur 208016, India 123 Proc. Natl. Acad. Sci., India, Sect. A Phys. Sci. DOI 10.1007/s40010-016-0296-7

Upload: others

Post on 01-Jul-2022

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Acceptorless Alcohol Dehydrogenation: A Mechanistic

REVIEW

Acceptorless Alcohol Dehydrogenation: A MechanisticPerspective

Pragati Pandey1 • Indranil Dutta1 • Jitendra K. Bera1

Received: 1 August 2016 /Accepted: 18 August 2016

� The National Academy of Sciences, India 2016

Abstract Alcohols are unreactive and require strong

inorganic oxidants to convert to synthetically useful car-

bonyl compounds. Acceptorless dehydrogenation of alco-

hol is a green and atom-economic alternative, which

provides aldehyde (or ketone) without the use of sacrificial

acceptor molecules and the side product is molecular

hydrogen. This review provides a brief overview of the

initial work followed by recent advances in the field of

acceptorless alcohol dehydrogenation. Catalysts that

employ metal–ligand cooperation for alcohol activation

and dehydrogenation are covered in details. Different

mechanisms are examined and clear advantages associated

with a bifunctional pathway are outlined. Mechanistic

understanding at the molecular level helps to develop new

generation dehydrogenation catalysts. Recent works from

our group on this area along with literature reports are

discussed.

Keywords Acceptorless alcohol dehydrogenation �Bifunctional catalysis � Metal–ligand cooperation �Dehydrogenation mechanism �Bifunctional double hydrogen transfer

Abbreviations

A Angstrom

AAD Acceptorless alcohol dehydrogenation

AD Alcohol dehydrogenation

ADHC Acceptorless dehydrogenative coupling

Bn Benzyl

bMepi 1,3-Bis(6-methyl-2-pyridylimino)isoindolate

KOtBu Potassium tert-butoxide

BDHT Bifunctional double hydrogen transfer

bpyO a,a0-BipyridonateCp Cyclopentadienyl

Cp* Pentamethylcyclopentadienyl

Cy Cyclohexyl

DABCO 1,4-Diazabicyclo[2.2.2]octane

DBU 1,8-Diazabicyclo[5.4.0]undec-7-ene

dppf 1,10-Bis(diphenylphosphino)ferroceneDFT Density functional theory

eu Energy unit

Et Ethyl

GC–MS Gas chromatography–mass spectrometry

HMB Hexamethyl benzene

IMes 1,3-Bis(2,4,6-trimethylphenyl)imidazol-2-

ylidene

IiPr 1,3-Diisopropyl imidazol-2-ylideneiPr Isopropyl

MLC Metal ligand cooperation

Me MethylnBu n-Butyl

NHC N-Heterocyclic carbine

NTs N-Tosyl

Nu Nucleophile

OAc Acetate

OTf Trifluoromethanesulphonate

Ph Phenyl

phenO 2,9-Dihydroxy-1,10-phenanthrolinetBu Tertiarybutyl

TOF Turn-over-frequency

TEMPO 2,2,6,6-Tetramethylpiperidine-1-oxyl

Tp0 Tris(3,5-dimethylpyrazolyl)borate

& Jitendra K. Bera

[email protected]

1 Department of Chemistry, Center for Environmental Sciences

and Engineering, Indian Institute of Technology Kanpur,

Kanpur 208016, India

123

Proc. Natl. Acad. Sci., India, Sect. A Phys. Sci.

DOI 10.1007/s40010-016-0296-7

Page 2: Acceptorless Alcohol Dehydrogenation: A Mechanistic

1 Introduction

X–H (X=C, N or O) bonds are abundant in organic mole-

cules. Removal of two hydrogens from adjacent atoms

provides an access to unsaturated molecules (Scheme 1)

[1]. For example, dehydrogenation of alkanes affords

alkenes which are powerful precursors for a diverse array

of useful products [2–8].

A positive aspect of this reaction is the possibility of

removal of hydrogens as molecular hydrogen. The desat-

uration mechanism in natural system proceeds via an initial

hydrogen atom abstraction by a metal-oxo species [9–11].

This is followed by another fast hydrogen atom transfer to

produce an olefin and the reduced form of the catalyst

(Scheme 2). The catalyst is then regenerated by a terminal

oxidant. It should be noted here that enzymatic desatura-

tion does not produce molecular hydrogen. However,

synthetic organometallic catalysts can dehydrogenate sat-

urated organic molecule where hydrogens are liberated as

molecular hydrogen. This short review describes recent

advances in the acceptorless dehydrogenation of alcohol.

Particular emphasis is placed on the mechanism of the

desaturation process. This is not a comprehensive review of

the literature. Rather, it highlights key developments in the

field from a mechanistic perspective.

Alcohols are common precursors in chemical reactions

although they are unreactive because of the hydroxy unit

(OH), which is a poor leaving group and is difficult to

displace [12]. Activation of alcohols is usually achieved by

turning the hydroxy into a better leaving group. Conven-

tional methods are either by protonating the hydroxy group

or by converting it into a sulfonate or a halide [13].

However, these activation methods have several short-

comings. In conventional organic transformations, the

oxidation/dehydrogenation of alcohols involve the utiliza-

tion of stoichiometric or excess amounts of inorganic

oxidants such as chromium(IV) reagents, peroxides or

pressurized oxygen which are hazardous [14–16].

Employment of various co-catalysts, additives and com-

bined catalytic systems of metal complexes and terminal

oxidants (such as TEMPO) yields undesirable stoichio-

metric waste (Scheme 3) [17, 18].

From an environmental viewpoint, there is a need to

develop synthetic protocols that have high atom-economy,

employ cheap and safe reagents and produce no hazardous

waste. Towards this goal, dehydrogenation methodology

without the use of conventional oxidants are developed

paving the way for the advancement of acceptorless alco-

hol dehydrogenation (AAD) (Scheme 4).

2 Acceptorless Alcohol Dehydrogenation (AAD)

AAD is essentially a reaction that removes one hydrogen

molecule from ubiquitous yet considerably less reactive

alcohols to form carbonyls—a more potent synthon. Lib-

eration of hydrogen without the use of stoichiometric

oxidant/acceptor makes the AAD a green and environ-

mentally benign synthetic methodology [19–26]. Extrusion

of hydrogen atoms from neighboring atoms of an organic

molecule in the form of molecular hydrogen is, in most

cases, a thermodynamically uphill process. To drive the

equilibrium towards the product generation, the molecular

hydrogen should be effectively removed from the reaction

mixture. Alternatively, the unsaturated intermediates gen-

erated during the reaction can also be hydrogenated with

in situ generated hydrogen, a process known as ‘borrowing

hydrogen’ method (Scheme 5) [27–29].

3 A Brief Overview

There have been considerable activities during past three

decades or so on various homogeneous catalysts for alcohol

dehydrogenation (AD), i.e. the conversion of alcohols to

aldehydes/ketones through the direct formation of molec-

ular hydrogen [30]. One of the earliest finding from

Scheme 1 Dehydrogenation of alkane

Scheme 2 Enzymatic desaturation of alkane

Scheme 3 Conventional routes for oxidation of substrate with

concomitant formation of stoichiometric waste

Scheme 4 Acceptorless dehydrogenation

P. Pandey et al.

123

Page 3: Acceptorless Alcohol Dehydrogenation: A Mechanistic

Dobson and Robinson group involved [Ru(OCOCF3)2(-

CO)(PPh3)2] (1) in the dehydrogenation of a series of pri-

mary and secondary alcohols [31]. This reaction required

an excess of a fluorinated carboxylic acid and was termed

as ‘acid-promoted’ reaction (Scheme 6). Jung and Garrou

introduced a class of parallel catalysts based on bidentate

diphosphine ligands and reported their mechanistic aspects

for dehydrogenation of primary alcohols [32].

Later Hulshof et al. found out that the dehydrogenation

of alcohols is often complicated due to decarbonylation

(reason of catalyst poisoning) and aldol condensations

under the reaction conditions [33]. Hulshof carried out their

studies with ruthenium complexes 2a–d containing fluori-

nated acid derivatives as ligands for the dehydrogenation of

alcohols (Scheme 7) [34].

Morton et al. [35, 36] introduced base promoted dehy-

drogenation catalyzed by various ruthenium and rhodium

complexes. It was the very first-time when alcohols of

lower molecular weight viz. ethanol, propanol and iso-

propanol were dehydrogenated with reasonable efficiency.

The activity for catalysts [RuH2(N2)(PPh3)3] and [RuH2(-

PPh3)4] could be significantly improved when the reactions

were carried out under visible light. Rhodium based cata-

lysts [RhCl(PPh3)3], [RhH(PiPr3)3], [RhCl((P(OPh)3)3]

were also employed for alcohol dehydrogenation and

resulted in moderate conversions [37, 38]. Wilkinson’s

catalyst was found inactive towards hydrogen generation

from isopropanol, however, addition of triethylamine

improved its activity [39, 40]. Under basic conditions, a

variety of ruthenium–arene and carbene complexes,

including [RuCl(PPh3)2Cp], [RuCl(PPh3)2(indenyl)],

[RuCl2(benzene)]2, [RuCl2(p–cymene)]2, [PhCH =

Ru(PCy3)2Cl2] and [Ru(IMes)–(PPh3)2CO(H)2], were suc-

cessfully tested for dehydrogenation of 1-phenylethanol

[41]. Beller and coworkers screened multiple ruthenium

precursors for dehydrogenation of isopropanol and found

[RuCl3.xH2O] and [RuCl2(p–cymene)]2 gave best TOF

when the reaction was carried out in the presence of two

equivalents of PCy3 [42]. In a subsequent study for iso-

propanol dehydrogenation, series of nitrogen donor ligands

were tested using the [RuCl2(p–cymene)]2 as precursor

[43]. Following this work, Albrecht, Madsen and Szym-

czak groups reported several catalysts based on various

phosphine, mesoionic triazolylidene, N-heterocyclic car-

bene and N,N,N–bMepi (bMepi = 1,3-bis(6-methyl-2-

pyridylimino)isoindolate) ligands which effectively per-

formed AAD reactions (Scheme 8) [44–54].

4 Classical Mechanism

Transition metal catalysts aid in the AAD reaction. Careful

mechanistic investigations reveal that classical AAD pro-

ceeds via an initial oxidative addition followed by b–Helimination of the metal-alkoxide leading to a metal-di-

hydride intermediate and the carbonyl product (Scheme 9)

[55–60]. The metal catalyst is then regenerated by the

liberation of a dihydrogen molecule via reductive elimi-

nation. The intermediacy of a metal-dihydride species is

confirmed by deutrated studies. When reactions are per-

formed with deuterated alcohol, a high level of H/D

scrambling in the final product supports a ‘dihydride

mechanism’ [61, 62]. This classical mechanism is associ-

ated with a number of disadvantages and limitations. Under

basic media, product aldehyde undergoes aldol-type rear-

rangement rendering the product separation very difficult.

Moreover, for certain metals like ruthenium, the metal

alkoxide is so stable that further b–hydride elimination is

hindered [63]. Also, redox adjustment of metal during the

catalytic cycle makes this an energetically unfavorable

process and usually requires elevated temperature to occur.

Hence, from synthetic and selectivity viewpoints, an

alternative approach was necessary which is devoid of such

shortcomings. This has been achieved by introduction of

bifunctional catalysts whose working principle is different

from a classical catalyst. The design principle of a

bifunctional catalyst and its mechanistic implications are

discussed in the following sections.

Scheme 5 Borrowing hydrogen methodology

Scheme 6 Acid-promoted dehydrogenation of alcohol

Acceptorless Alcohol Dehydrogenation: A Mechanistic Perspective

123

Page 4: Acceptorless Alcohol Dehydrogenation: A Mechanistic

Scheme 7 Dimeric ruthenium catalysts containing acid derivatives for hydrogen generation

Ru ClN

N N

PhRu ClN

NN

N

OTf

Ru ClNN

N

O

O

MesRu Cl

NN N N

Ph

PF6

Ru ClNN

N

OSitBuMe2

Cl

Mes

Ru ClNN

NCl

Mes

Ru

Cl

NNN

N O

N

+Ph

NN

NIr

NIr Cp*

Cl

Cl

Cp*

NOTf

Ru ClN

NX

R1

R2R1 = ipr, CH2Ph, cyR2 = ipr, CH2Ph, cyX = Cl, PPh3n = 0, 1

n+

NN

NN

NRu

PPh3

PPh3

H

3 4 5 6

7 8 9

10 1211

Scheme 8 List of transition metal complexes used for AAD reaction

Scheme 9 Classical

mechanism of alcohol

dehydrogenation

P. Pandey et al.

123

Page 5: Acceptorless Alcohol Dehydrogenation: A Mechanistic

5 Cooperative Catalysis

Traditional concept of a catalyst is based on the fact that

during catalysis, the metal center is the active site whereas

ligands are only present to provide steric and electronic

modulation. However, a new concept has emerged where

synergic participation of two or more chemical function-

alities carries out a difficult chemical process [64–68]. Both

metal–metal and metal–ligand cooperation (MLC) strate-

gies have been implemented in synthetic catalysts. Coop-

erative participation of two metals in close proximity has

been demonstrated to activate small molecule, controls

stereo-electronic features of a chemical process and facil-

itates product elimination (Scheme 10) [69–72]. Synthetic

catalysts are also designed where a cooperating ligand at

the vicinity of active metal site actively participates in

bond-activation process and undergoes reversible chemical

transformation during the catalytic cycle to make the pro-

cess more efficient, selective and atom-economic

(Scheme 11) [73–77]. The MLC has emerged as one of the

powerful concept to develop organometallic catalysts

[78–82].

This MLC paradigm offers several distinct advantages

for bond activation chemistry—(a) the combination of a

Lewis acid (metal ion) and Lewis base (ligand) pair is

particularly effective for polarizing neutral molecule

(e.g., H2); (b) the oxidative addition/reductive elimina-

tion steps are not required which, in principle, allow 3d

metals for small molecule activation chemistry; [83–85]

(c) a bifunctional mechanism affords thermodynamic

favorability for the activation of substrates. This strategy

has been used in Bera’s group for small molecule acti-

vation [86, 87].

6 Design Strategies of Bifunctional Catalysts

The design strategy of a bifunctional catalyst involves a

metal unit and a conjugated proton-responsive arm posi-

tioned at an appropriate place on the ligand architecture.

The synergistic interplay between the Lewis acid metal and

the Lewis base ligand facilitates substrate recognition,

activation, and transformation [88–94]. Efficiency and

selectivity of the catalysis can be tuned via electronic

diversity and structural flexibility of the cooperating

ligands. Proton responsive groups such as NH and OH

functional units are ideal targets because of their easy

accessibility, chemical stability and synthetic feasibility.

These groups serve several purposes viz. coordinating

groups, hydrogen bonding donors, hydrogen bonding

acceptors, and/or proton sources. The position of these

proton responsive groups is also vital in the catalyst design.

Depending upon the positioning of these groups with

respect to the metal, they can be classified as a, b and cprotic functionalized MLC. AAD reactions that occur

through a, b and c protic complexes are discussed below.

6.1 a–Protic Bifunctional Catalyst

There is a range of catalytic systems that employ a–proticamine/amido complexes and AAD and related reactions.

The cooperating group of the ligand in this category (i.e. aposition) binds to the metal directly. The most celebrated

example is the hydrogenation catalysts by Noyori and co-

workers [95–99]. Noyori system 13 is an example of abifunctionality which displays a chelate-stabilized inter-

convertible amido/amine couple (Scheme 12).

Another interesting example in this category is Grutz-

macher’s catalyst 14 that exploits the interaction between

Lewis acidic Rh and Lewis basic amide nitrogen

(Scheme 13) [100–103]. Although it requires cyclohex-

anone or methylmethacrylate as hydrogen acceptor, its

catalytic efficiency has been discussed for efficient dehy-

drogenative coupling of primary alcohols with amine,

water or methanol. Low catalyst loading, good chemose-

lectivities and high functional-group tolerance were found

in the reaction signifying its utility in the synthetic process.

Scheme 10 Bond activation via metal–metal cooperation

Scheme 11 Bond activation via metal–ligand cooperation

Scheme 12 Noyori’s a-protic amine/amido hydrogenation catalyst

Acceptorless Alcohol Dehydrogenation: A Mechanistic Perspective

123

Page 6: Acceptorless Alcohol Dehydrogenation: A Mechanistic

Recently, Baratta et al. [104] have reported Ru and Os

based a–protic amine/amido system. These were employed

to catalyze AD of secondary alcohols to ketones

(Scheme 14). Ru complex 15 showed greater catalytic

efficiency than the corresponding Os complex 16 for AD.

Beller and co-workers introduced a–protic NH func-

tionalized Ru–PNP pincer complexes for dehydrogenation

of ethanol [105]. Complex 17 and 18 with NH functionality

show superior catalytic activity than complex 19 which is

devoid of NH proton at a position (Scheme 15). The active

form of the Ru–amido catalyst is achieved by treating the

complex 17 with NaOEt followed by elimination of

molecular hydrogen (Scheme 16).

Beller et al. has also reported 3d metal-based Fe com-

plexes 20 and 21 (Scheme 17) with aliphatic PNP ligand

which were found to be catalytically active in AAD under

base-free conditions [106]. It was interesting to note that

catalyst 20 was found to dehydrogenate methanol in the

presence of KOH (Scheme 18). Later on complex 20 and

21 were further investigated and expanded by Jones and

Schneider for AAD reactions (Scheme 19) [107].

Compared to metal–amide/amine complexes, the use of

metal–alkoxide/alcohol complexes for MLC is less com-

mon. This is attributed to the reduced basicity of coordi-

nated alkoxides and higher lability of coordinated alcohol

in late transition metal complexes. Gelman developed a

dibenzobarrelene-based PCsp3P pincer ligand and synthe-

sized an iridium complex 22 [108–110]. This contains a

polar alcohol/alkoxide based CH2–OH hemilabile sidearm

at a position to the metal center. This catalyst showed very

high catalytic efficiency in the AD of primary and sec-

ondary alcohols (Scheme 20).

Scheme 13 Grutzmacher’s catalyst for alcohol dehydrogenation in

presence of acceptor

M

H2N

Ph2P

Cl

Cl

NH2

PPh2

Fe

OH

R'R

O

R R'0.4 mol %15-16

KOtBu, 130 °C

15: M = Ru16: M = Os

H2

Scheme 14 AD of alcohols catalyzed by ruthenium and osmium

based a-protic amine/amido system

Scheme 15 Crucial role of NH functionality in the dehydrogenation

of ethanol

Scheme 16 Ruthenium-catalyzed dehydrogenation of ethanol to

ethyl acetate

Scheme 17 Iron based aliphatic PNP pincer complexes for AAD

Scheme 18 Iron-catalyzed methanol dehydrogenation

P. Pandey et al.

123

Page 7: Acceptorless Alcohol Dehydrogenation: A Mechanistic

Recently Jones and coworkers have reported a nickel(II)

complex 23, supported by tris(3,5-dimethylpyrazolyl)bo-

rate ligand and 2-hydroxyquinoline ancillary ligand effec-

tively catalyzing AD of a variety of alcohols to afford

ketones, esters and lactones (Scheme 21) [111]. The pres-

ence of -OH group at the ortho position of quinoline is

crucial for AAD reaction as in the presence of 8-hydrox-

yquinoline, reaction doesn’t occur. Bera group has also

synthesized metal–metal singly bonded diruthenium com-

plexes, featuring a hydroxy appendage for AD reactions.

Detailed reactivity and mechanistic studies are discussed

later.

6.2 b–Protic Bifunctional Catalyst

b–Protic systems are also known where the NH/OH moi-

eties are not directly bonded to the metal center. Shvo’s

catalyst 25 is a landmark example of bifunctional catalyst

where the cooperative sites are built in the cyclopentadi-

enyl ring and positioned b to the metal [112–115].

Although the original catalyst is a dinuclear complex 24,

the active species contains a single Ru metal center

(Scheme 22) [116]. Shvo’s catalyst dehydrogenates sec-

ondary alcohols to the corresponding ketones. A concerted

migration of hydride and proton to the metal and ligand

center respectively is involved in this process. The bond

activation by MLC in Shvo’s system engages intercon-

version between Ru complexes with the g5–bound

Scheme 19 Proposed mechanisms for iron-catalyzed AAD

Scheme 20 a Elimination of H2 from complex 22. b AAD mech-

anism by complex 22

Scheme 21 Proposed mechanism for the nickel-catalyzed AAD

Acceptorless Alcohol Dehydrogenation: A Mechanistic Perspective

123

Page 8: Acceptorless Alcohol Dehydrogenation: A Mechanistic

hydroxycyclopentadienyl and the g4–bound cyclopenta-

dienone moieties (Scheme 23).

6.3 c–Protic Bifunctional Catalyst: Aromatization/

Dearomatization

So far we have discussed systems where proton responsive

groups are disposed at a or b position relative to the metal

center and MLC operates via protonation/deprotonation.

Milstein group introduced systems with long distance c–protic functionality [21, 72, 117]. In the course of MLC,

ligand aromatic skeleton is disrupted and restored (arom-

atization/dearomatization) during the bond breaking/for-

mation process creating a platform for catalysis. A large

number of tridentate pincer ligands have been designed

which follow this strategy. The substituted lutidines or

2-picolines having one or two CH2 moiety in the ortho

position(s) of the central pyridine constitute the main

skeleton. This on deprotonation, by treatment with strong

base, undergoes dearomatization at the heteroaromatic unit

with generation of an exocyclic double bond and thus

resulting in active centers for MLC (Scheme 24). During

this process the pyridine ring loses its aromaticity and

N-atom act as amide donor. Alcohol is added to it restoring

the aromaticity. A metal-dihydride is generated via b-hy-dride elimination of the metal-alkoxide and aldehyde is

produced. The PNP- and PNN-type ruthenium pincer cat-

alysts (26 and 27) were found highly active for dehydro-

genative synthesis of imines, amides, esters and acetals

from alcohols [118–123].

However, in all these reactions, catalytic amount of base

is necessary for catalyst activation. Recently Milstein

group has synthesized electron-rich PNP and PNN type

ruthenium(II) hydrido borohydride pincer complexes (28,

29) which can catalyze alcohol dehydrogenation reactions

in the absence of a base [124]. The superior activity of

PNN complex compared to PNP complex is attributed to

the fact that PNN complex contains a heamilabile group

which provides a vacant site during catalysis (Scheme 25).

Using a similar protocol, several Cp*Ir based complexes

(30–32) have been synthesized by Yamaguchi, Fujita, and

Tanino using 2-hydroxypyridine, a,a0-bipyridonate (bpyO),and 2,9-dihydroxy-1,10-phenanthroline (phenO) as ligands,

respectively (Scheme 26) [125–129]. These catalysts are

proven to be an efficient platform for the AAD of alcohols

utilizing the concept of lactam–lactim tautomerism

(Scheme 27). It has been observed that catalyst 31

Cp*Ir(bpyO)(H2O) with ligand bpyO is more reactive than

30 with 2-hydroxypyridine as ligand or catalyst 32

Cp*Ir(phenO)(H2O) with phenO ligand [130]. The differ-

ence in reactivity can be rationalized as complex 31 is the

active form of the catalyst whereas in 30, hydroxypyridine

unit must be deprotonated to generate active species. The

decreased catalytic activity of 32 is attributed to the

reduced proton affinity of oxygen atom of phenO ligand

and to the lower ‘aromatization effect’ in highly conjugated

phenO ligand. Further, Yamaguchi and coworkers have

also developed the Rh and Ru complexes 33

Cp*Rh(bpyO)(H2O) and 34 (HMB)Ru(bpyO)(H2O) with

bypO ligand (Scheme 26) [131]. These catalysts have also

shown promising results for AAD reactions.

7 Mechanism of Bifunctional Catalysts

The AAD mechanism by a classical catalyst proceeds

through the intermediacy of a metal-dihydride intermediate

generated via alcohol oxidative addition followed by b–Helimination, and the catalyst is regenerated by reductive

elimination. But bifunctional catalyst involves transfer of a

a-C–H of alcohol to the metal center and proton of –OH

group to a ligand heteroatom thus generating a metal–

monohydride [61, 112]. Depending upon the nature of the

catalysts, there are two possible pathways that lead to the

formation of metal–monohydride intermediate during

catalysis—(a) stepwise inner-sphere mechanism via b–Helimination; (b) concerted transfer of proton and hydride

Scheme 22 Dimeric Shvo’s

complex and its active form in

solution

Scheme 23 Proposed mechanistic pathway for hydrogen transfer

involving alcohols using Shvo’s complex

P. Pandey et al.

123

Page 9: Acceptorless Alcohol Dehydrogenation: A Mechanistic

via outer-sphere bifunctional double hydrogen transfer

(BDHT) (Scheme 28) [53, 127, 132, 133].

Ligand assisted inner-sphere mechanism involves direct

binding of substrate to the metal center. Such binding

necessarily requires a vacant site at the metal center, and

hence this is thermodynamically less favorable. However,

once metal–alkoxide bond formation takes place, it is

energetically feasible and driven towards product forma-

tion. However, metal–alkoxide bond being very stable, the

catalyst regeneration is sometimes kinetically not permitted

and it often leads to the catalyst deactivation after few

catalytic cycles. Extensive kinetic, deuterated and DFT

studies have been undertaken to decipher the mechanism

involved [49, 53, 134–136]. These studies revealed that the

rate limiting step is invariably the b–H elimination step.

Moreover, temperature dependence studies using Arrhenius

and Eyring plots give estimates of the activation energy of

the processes, which in turn predict if the reaction is

thermodynamically favored or forbidden. In literature, a

range of DS� values (?12 to -30 eu) are estimated for a b–H elimination turnover-limiting step from metal–alkoxide

species [137, 138]. Also as discussed, the metal–alkoxide

intermediate is a stable species and its direct isolation and

characterization is often the most convenient way to

understand the mechanism.

Contrary to that, in outer-sphere mechanism, substrate

does not bind directly to the metal center. Rather, syner-

gistic interaction of the catalyst (metal and ligand) with

alcohol leads to the product formation. Migration of proton

and hydride to the ligand and the metal center, respec-

tively, is a concerted process and proceeds via a six-

member transition state. This is followed by elimination of

molecular hydrogen and the catalyst regeneration. Such

type of proton–hydride shuttle between ligand and metal

center signifies the necessity of proton responsive ligand

arm such as NH/OH group in bifunctional catalysts.

Although the idea of BDHT is simple and uncomplicated,

designing experiments to conclusively establish this

Scheme 24 Aromatization–dearomatization during alcohol activation

Scheme 25 Dehydrogenation of secondary alcohol by 28, 29 in

absence of base

Scheme 26 A series of Cp*Ir,

Rh, and Ru complexes

developed by Yamaguchi and

co-workers for the AAD

reaction

Acceptorless Alcohol Dehydrogenation: A Mechanistic Perspective

123

Page 10: Acceptorless Alcohol Dehydrogenation: A Mechanistic

mechanism is a challenging task. Kinetic studies unravel

the rate and the order of the reaction and Hammett plots are

also helpful in understanding the reaction mechanism.

Temperature dependence studies also give valuable infor-

mation regarding the activation parameters. Contrary to an

inner-sphere mechanism, relatively large negative entropy

of activation DS� values (*-30 eu) are indicative of a

BDHT mechanism [139]. This not only supports an asso-

ciative process but also suggests a higher degree of orga-

nization in the transition state, indicating that the catalyst

and the substrate are involved in association prior to

migration of proton and hydride. Analyzing deuterium

content in the final products while using deuterated sub-

strate give a comprehensive picture of the mechanism. DFT

calculations have also become an important tool particu-

larly for proposing the possible intermediates and the

energetics involved during the course of the reaction [140].

8 Metal–Metal Bonded Platform

All bifunctional catalytic systems mentioned above utilize

a single metal ion. The utility of metal–metal cooperation

in organometallic catalysis has been demonstrated on sin-

gly-bonded [Ru–Ru] systems [141]. In general, b-hydrideelimination of a metal–alkoxide intermediate occurs on a

single metal centre and proceeds via a four-membered

agostic species [142]. A bimetallic construct provides a

scope for cooperative b-hydride elimination that involves

both metals and may turn out to be energetically more

favourable (Scheme 29) [143].

Towards this objective, a crescent shaped 1,8-naph-

thyridine-diimine ligand is synthesized. Treatment of this

ligand with Ru2(OAc)4Cl resulted in the formation of a

metal–metal bonded compound 35. The catalytic efficacy

of 35 was examined for AAD reactions for a range of

Scheme 27 Proposed mechanism of AAD employing lactam–lactim tautomerism

Scheme 28 Generalized

mechanistic pathway for

a inner-sphere and b outer-

sphere mechanism

P. Pandey et al.

123

Page 11: Acceptorless Alcohol Dehydrogenation: A Mechanistic

alcohols. Catalyst 35 (1 mol%) afforded 89 % conversion

of benzyl alcohol to benzaldehyde at 70 �C in toluene for

6 h. Optimization studies showed that KOH was the best

among a variety of bases. This was further extended for

acceptorless dehydrogenative coupling (ADHC) reactions.

A mixture of benzyl alcohol, triphenylphosphonium

methoxycarbonylmethylide (Wittig reagent, 1.5 equiva-

lents), 1 mol% 35 and 10 mol% KOH afforded E-methyl

cinnamate predominantly as confirmed by NMR analysis

(Scheme 30).

Carrying out AD reaction of benzyl alcohol under

identical reaction conditions with catalyst Ru2(OAc)4Cl,

having accessible axial site, yielded only 30 % of product.

This suggests that mere presence of a vacant axial site

around metal is not the sole requirement for product for-

mation. Rather, suitably designed ligand framework renders

trans ligands labile providing equatorial sites accessible for

catalysis. Accordingly, a mechanism has been proposed

where proceeds on the equatorial platform (Scheme 31).

Initially, the acetate group trans to the naphthyridine unit is

replaced by the alkoxide moiety. This is followed by a

bimetallic b-hydride elimination resulting in a Ru-hydride

intermediate along with the formation of aldehyde. The

aldehyde is then extruded and an alcohol molecule binds to

the metal. The catalyst is regenerated via a dehydrogenation

pathway involving an intramolecular proton transfer from

alcohol to the metal–bound hydride. Exhaustive kinetic

studies favored the proposed bimetallic mechanism. To find

out the order of the reaction with respect to catalyst 35,

initial rate was monitored. The linear relationship between

initial rate and the catalyst concentration ascertained that

the reaction is first-order with respect to 35 (Fig. 1a).

Scheme 29 Agostic interaction during b-hydride elimination on

a monometallic and b bimetallic platform

Scheme 30 Acceptorless dehydrogenative coupling using Wittig

reagent

Scheme 31 Proposed

mechanistic pathway for AAD

reaction

Acceptorless Alcohol Dehydrogenation: A Mechanistic Perspective

123

Page 12: Acceptorless Alcohol Dehydrogenation: A Mechanistic

Using integrated rate law for the reaction of the type

A ? B with the constraint [A] = 1 and [B] = 0 at t = 0,

the ln[A] vs time plot shows a first-order kinetics (Fig. 1b).

These experimental findings taken together suggested that

one molecule of the catalyst 35 and alcohol are involved in

the rate-determining step. As one molecule of the catalyst

35 consists of two ruthenium centers, it is logical to pre-

sume that the reaction occurs on the bimetallic assembly.

Deuterium lebling study using a,a-[D2]-benzyl alcohol

showed deuterated benzaldehyde as the major product

(92:8 D/H, observed by GC–MS analysis) indicating the

involvement of a cooperative mechanism (Scheme 32).

A comparative study of the two reactions, (a) PhCH2OH

in toluene and (b) PhCD2OH in toluene–d8 showed kC–H/

kC–D = 2.71 ± 0.04 (Fig. 2). This demonstrated that the

C–H bond-breaking is one of the slower steps in the

reaction. When PhCH2OD was used as a substrate instead

of PhCH2OH, the rate of reaction is 4.94 ± 0.02 times

slower (Fig. 2). High kO–H/kO–D value is indicative of the

fact that elimination of molecular hydrogen during the final

stage of the catalytic cycle is likely to be the rate limiting

step. This proposition was further supported by the DFT

calculations which revealed that the dehydrogenation step

is most exothermic in nature (14.64 kcal/mol) (Fig. 3).

Scrutiny of a number of literature reports revealed that

AAD of primary alcohols by bifunctional catalysts invari-

ably yields esters as major products. Ester formation

process is rationalized by invoking Tischenko type reaction

or by hemiacetalyzation followed by dehydrogenation

[144–147]. The observed aldehyde selectivity in this

reaction was explained on the basis that aldehyde must

binds to the metal centre for effective hemiacetalyzation.

But the ligand architecture ensures that the aldehyde is

rapidly extruded from the [Ru = Ru] core nullifying any

possibility of hemiacetalyzation.

Another interesting aspect of this reaction is that the b-hydride elimination step proceeds via a five-membered

transition state involving two metal centers. The energy

requirement is lesser compared to a system where b-hy-dride elimination happens on a single metal center

involving a four-membered transition state. As a result, this

is a unique example where dehydrogenation is more

energy-demanding than b-hydride elimination indicating

metal–metal cooperation is operative here.

9 Dual Metal–Metal and Metal–LigandCooperation: Selective Synthesis of Imine

A bifunctional catalyst is developed on a metal–metal

bonded platform which displays both metal–metal and

metal–ligand cooperativity. The design strategy involved

the introduction of a protonic arm (–OH) on [Ru2(CO)4]2?

platform utilizing a naphthyridine functionalized N-hete-

rocyclic carbene (NHC) ligand [148]. This hydroxy unit,

positioned at site trans to the metal–metal bond, plays a

crucial role in exhibiting metal–ligand cooperation.

(Scheme 33).

A range of catalysts (36–39) were synthesized via aldol-

type C–C bond formation reactions using a variety of

electron-deficient aromatic aldehydes [149]. The diruthe-

nium core is bridged by the ligand where NHC unit binds

Fig. 1 a Dependence of initial rate on 35 and b Decay of benzyl alcohol vs time. Reprinted with permission from ref 143. Copyright 2016

American Chemical Society

OH

DD O

D/H

1 mol% 45

10 mol% KOHToluene, 70 °C

D/H = 92:8

Scheme 32 Deuterium scrambling using a,a-[D2]-benzyl alcohol for

AAD reaction

P. Pandey et al.

123

Page 13: Acceptorless Alcohol Dehydrogenation: A Mechanistic

one axial site and the other site is occupied by the hydroxy

arm (Scheme 34). Catalyst 36 (1 mol%) afforded 98 %

conversion of benzyl alcohol to benzaldehyde under reflux

in toluene for 24 h. This has further been extended for

ADHC reactions. When 1.2 mmol of benzylamine was

added in presence of 4 A molecular sieves, N-benzylidine

benzylamine formed selectively (Scheme 35).

This reaction was found to be effective with a variety of

bases (DBU, DABCO, KOH, KOtBu, NaH). However,

further studies were carried with DABCO. With different

combinations of alcohols and amines, a set of total 25

reactions was carried out. The yields varied in the range

71–96 %. In order to understand the role of the hydroxy

appendage in 36, a comparative study with a similar

complex but devoid of the hydroxy side arm was also

undertaken. Catalyst 40 afforded dehydrogenation product

benzaldehyde in significantly lower yield (55 %)

(Scheme 36).

Similarly, corresponding imine conversion was found to

be much less. These observations clearly illustrate the

crucial role of the hydroxy appendage for the catalyst

activity. Accordingly, a mechanism is proposed using the

concept of bifunctionality. DFT calculations have been

carried out and the computed energetics support the

proposition. The active catalyst 41 is the deprotonated form

of 36, obtained on treatment with base. Employment of 41

as catalyst leads to product formation under base-free

condition. At first, alcohol is activated in a bifunctional

fashion by 41 to give [Ru–Ru]–alkoxide(axial) and the

hydroxy arm is opened up (Scheme 37). b-hydride elimi-

nation of the [Ru–Ru]-alkoxide affords aldehyde and a

[Ru–Ru]–H intermediate is generated, which is identified

by 1H NMR spectrum having a characteristic signal at

d = -7.37 ppm). Elimination of molecular hydrogen

generates the active catalyst.

The extruded aldehyde reacts with amine to give imine

as the final product. To gain further insight, kinetic Ham-

mett studies for key b-hydride elimination step was carried

out. A plot of ln(c0/c) of substituted benzyl alcohols against

the same values for benzyl alcohol resulted in a linear plot

confirming first order dependence of alcohol. The slopes of

the straight lines were plotted against all possible r values

(r?, r-, r.) of a particular substituent and a linear rela-

tionship is only obtained only with r? (Fig. 4).

This supports our proposition that a positive charge is

generated at the benzylic position of alcohol during b-hy-dride elimination. The possibility of a bifunctional mech-

anism over the classical mechanism, which necessarily

involves oxidative addition of alcohol to a low-valent

metal, has also been scrutinized by deuterated studies. As

discussed earlier, a classical mechanism of alcohol acti-

vation is associated with significant hydrogen scrambling

in the product imine. Madsen et al. employed catalyst

[RuCl2(IiPr)(p-cymene)] which lacks the metal–ligand

cooperation for AAD reactions. They observed 42 %

hydrogen incorporation in the final product when a,a–[D2]–

benzyl alcohol was used as substrate [49]. However, using

catalyst 36, a reaction of a,a–[D2]–benzyl alcohol with

Fig. 2 Reaction rates for PhCH2OH, PhCD2OH and PhCH2OD

versus time (min). Reprinted with permission from ref 143. Copyright

2016 American Chemical Society

Fig. 3 Computed reaction profile for AAD by catalyst 35. Energies

are shown in kcal/mol relative to A. Reprinted with permission from

ref 143. Copyright 2016 American Chemical Society

Scheme 33 Metal–ligand cooperation at axial site of a diruthenium

platform

Acceptorless Alcohol Dehydrogenation: A Mechanistic Perspective

123

Page 14: Acceptorless Alcohol Dehydrogenation: A Mechanistic

benzylamine resulted the formation of deuterated N-ben-

zylidene benzylamine as major product (93:7 D/H

observed by GC–MS analysis), supporting the proposed

bifunctional mechanism (Scheme 38).

Although the reaction occurs at axial site of the [Ru–Ru]

bond, bridging acetate plays an important role during the

catalysis. For the b-elimination to occur, an accessible site

at the metal centre is required. During the reaction, the

bridging acetate changes its coordination motif from l2 tog1 providing a vacant site.

The most important aspect of this reaction is the selec-

tive formation of imine. Reaction of amine and aldehyde

invariably gives imine when it is carried out without the aid

of a catalyst [150]. However, in a metal-catalyzed reaction,

Scheme 34 Syntheses of

diruthenium-NHC complexes

containing hemilabile protonic

arm at the axial site

Scheme 35 Selective imine formation via ADHC of alcohols with

amines catalyzed by 36

Scheme 36 Catalyst 40 devoid

of the hydroxy unit

Scheme 37 Proposed

mechanistic pathway for imine

formation

Fig. 4 Hammett study for the imination of alcohol. Reprinted with

permission from ref 148. Copyright 2014 by John Wiley & Sons, Inc

P. Pandey et al.

123

Page 15: Acceptorless Alcohol Dehydrogenation: A Mechanistic

there are two possibilities—(a) the amine attacks the metal-

coordinated aldehyde leading to a hemiaminal, which

subsequently undergoes dehydrogenation to generate

amide; (b) imine is formed by means of simple dehydra-

tion. The amine or N-alkylated products are also expected

as hydrogen is generated in the reaction. The ability of the

intermediate aldehyde to bind to the metal essentially

dictates the final product. For amide formation, aldehyde

must coordinate to the metal centre. However, the ligand

architecture in catalyst 36 makes it difficult for the alde-

hyde to bind the metal. Furthermore, the strong trans effect

of the axial NHC unit does not allow strong alcohol

binding. As a consequence, aldehyde is extruded from the

metal coordination sphere, which subsequently reacts with

amine to give imine. Use of molecular sieves to arrest

water molecules favors the reaction. Thus, this catalyst

exhibits both metal–ligand and metal–metal cooperations

for selective imine formation via dehydrogenative

coupling.

10 Future Outlook

Dehydrogenation is an important reaction that affords

unsaturated compounds and produce molecular hydrogen.

Alcohols are readily dehydrogenated to give carbonyl

compounds catalyzed by a variety of metal catalyst without

the use stoichiometric amount of oxidants or acceptor

molecules. The added advantage is the liberated hydrogen.

This raises the prospect of using these catalysts for rever-

sible dehydrogenation of organic liquid fuel for energy

storage and generation. Dehydrogenative coupling reac-

tions provide a vast array of C–C coupled products such as

imines, amines, amides and esters. Bifunctional catalysts

are particularly attractive since they activate alcohol

without redox change on the metal. Furthermore, metal–

ligand cooperation facilitates substrate activation and paves

a low-energy dehydrogenation pathway. Efforts are on to

develop bifunctional catalysts on diverse molecular plat-

forms involving different activation modes. Bimetallic

platform is shown to be effective for AAD reaction. Both

metal–metal and metal–ligand cooperations are exploited

to obtain selective dehydrogenative coupled products. A

clear understanding of AAD reactions would lead to cata-

lysts suited for more challenging amine and alkane

dehydrogenation reactions to obtain nitrile and alkenes

respectively.

Acknowledgments This work is financially supported by the

Department of Science and Technology (DST), India, and the Council

of Scientific and Industrial Research (CSIR), India. J.K.B. thanks

Department of Atomic Energy for DAE outstanding investigator

award. P.P and I.D. thank CSIR, India for fellowships.

References

1. Dobereiner GE, Crabtree RH (2010) Dehydrogenation as a

substrate-activating strategy in homogeneous transition-metal

catalysis. Chem Rev 110:681–703

2. West JG, Huang D, Sorensen EJ (2015) Acceptorless dehydro-

genation of small molecules through cooperative base metal

catalysis. Nat Commun 6:10093

3. Janowicz AH, Bergman RG (1982) C-H activation in com-

pletely saturated hydrocarbons: direct observation of M ? R–

H ? M(R)(H). J Am Chem Soc 104:352–354

4. Hoyano JK, Graham WAG (1982) Oxidative addition of car-

bon–hydrogen bonds of neopentane and cyclohexane to a pho-

tochemically generated iridium(I) complex. J Am Chem Soc

104:3723–3725

5. Beadry D, Ephritikhine M, Felkin H, Holmes-Smith R (1983)

The selective catalytic conversion of cycloalkanes into

cycloalkenes using a soluble rhenium polyhydride system.

J Chem Soc Chem Commun 788–789. doi:

10.1039/C39830000788

6. Maguire JA, Boese WT, Goldman AS (1989) Photochemical

dehydrogenation of alkanes catalyzed by trans-carbonylchloro-

bis(trimethylphosphine)rhodium: aspects of selectivity and

mechanism. J Am Chem Soc 111:7088–7093

7. Xu W, Rosini GP, Gupta M, Jensen CM, Kaska WC,

KroghJespersen K, Goldman AS (1997) Thermochemical alkane

dehydrogenation catalyzed in solution without the use of a

hydrogen acceptor. Chem Commun 2273–2274. doi:

10.1039/A705105K

8. Liu F, Goldman AS (1999) Efficient thermochemical alkane

dehydrogenation and isomerization catalyzed by an iridium

pincer complex. Chem Commun 655–656. doi:

10.1039/A900631A

9. Buist PH (2004) Fatty acid desaturase: selecting the dehydro-

genation channel. Nat Prod Rep 21:249–262

10. Breslow R, Baldwin S, Flechtner T, Kalicky P, Liu S, Washburn

W (1973) Remote oxidation of steroids by photolysis of attached

benzophenone groups. J Am Chem Soc 95:3251–3262

11. Bigi MA, Reed SA, White MC (2011) Diverting non-haem iron

catalysed aliphatic C–H hydroxylations towards desaturations.

Nat Chem 3:216–222

12. Salvatore RN, Yoon CH, Jung KW (2001) Synthesis of sec-

ondary amines. Tetrahedron 57:7785–7811

13. Reynolds DD, Kenyon WO (1950) Preparation and reactions of

sulfonic esters. III. Reaction of polyvinyl sulfonates with pri-

mary and secondary amines. J Am Chem Soc 72:1591–1593

Scheme 38 Deuterium

scrambling study of Imination

with benzyl alcohol-a,a-[D2] by

catalyst 36

Acceptorless Alcohol Dehydrogenation: A Mechanistic Perspective

123

Page 16: Acceptorless Alcohol Dehydrogenation: A Mechanistic

14. Muzart J (1992) Chromium-catalyzed oxidations in organic

synthesis. Chem Rev 92:113–140

15. Chen B, Wang L, Gao S (2015) Recent advances in aerobic

oxidation of alcohols and amines to imines. ACS Catal

5:5851–5876

16. Punniyamurthy T, Velusamy S, Iqbal J (2005) Recent advances

in transition metal catalyzed oxidation of organic substrates with

molecular oxygen. Chem Rev 105:2329–2364

17. Tojo G, Fernandez M (2007) Oxidation of alcohols to aldehydes

and ketones: a guide to current common practice. Springer, New

York

18. Ryland BL, Stahl SS (2014) Practical aerobic oxidations of

alcohols and amines with homogeneous copper/TEMPO and

related catalyst systems. Angew Chem Int Ed 53:8824–8838

19. Watson AJA, Williams JMJ (2010) The give and take of alcohol

activation. Science 329:635–636

20. Gunanathan C, Milstein D (2013) Applications of acceptorless

dehydrogenation and related transformations in chemical syn-

thesis. Science 341:1229712

21. Gunanathan C, Milstein D (2014) Bond activation and catalysis

by ruthenium pincer complexes. Chem Rev 114:12024–12087

22. Muthaiah S, Hong SH (2012) Acceptorless and base-free

dehydrogenation of alcohols and amines using ruthenium-hy-

dride complexes. Adv Synth Catal 354:3045–3053

23. Guillena G, Ramon DJ, Yus M (2010) Hydrogen autotransfer in

the N-alkylation of amines and related compounds using alco-

hols and amines as electrophiles. Chem Rev 110:1611–1641

24. Nielsen M, Kammer A, Cozzula D, Junge H, Gladiali S, Beller

M (2011) Efficient hydrogen production from alcohols under

mild reaction conditions. Angew Chem Int Ed 50:9593–9597

25. Shahane S, Fischmeister C, Bruneau C (2012) Acceptorless

ruthenium catalyzed dehydrogenation of alcohols to ketones and

esters. Catal Sci Technol 2:1425–1428

26. Bonitatibus PJ Jr, Chakrabortyb S, Dohertya MD, Siclovana O,

Jones WD, Soloveichika GL (2015) Reversible catalytic dehy-

drogenation of alcohols for energy storage. PNAS

112:1687–1692

27. Nixon TD, Whittlesey MK, Williams JMJ (2009) ChemInform

abstract: transition metal catalysed reactions of alcohols using

borrowing hydrogen methodology. Dalton Trans 753–762

28. Hamid MHSA, Slatford PA, Williams JMJ (2007) Borrowing

hydrogen in the activation of alcohols. Adv Synth Catal

349:1555–1575

29. Edwards MG, Jazzar RFR, Paine BM, Shermer DJ, Whittlesey

MK, Williams JMJ, Edney DD (2004) Borrowing hydrogen: a

catalytic route to C–C bond formation from alcohols. Chem

Commun 90–91. doi:10.1039/B312162C

30. Johnson TC, Morris DJ, Wills M (2010) Hydrogen generation

from formic acid and alcohols using homogeneous catalysts.

Chem Soc Rev 39:81–88

31. Dobson A, Robinson SD (1977) Complexes of the platinum

metals. 7. Homogeneous ruthenium and osmium catalysts for the

dehydrogenation of primary and secondary alcohols. Inorg

Chem 16:137–142

32. Jung CW, Garrou PE (1982) Dehydrogenation of alcohols and

hydrogenation of aldehydes using homogeneous ruthenium

catalysts. Organometallics 1:658–666

33. Ligthart GBWL, Meijer RH, Donners MPJ, Meuldijk J, Veke-

mans JAJM, Hulshof LA (2003) Highly sustainable catalytic

dehydrogenation of alcohols with evolution of hydrogen gas.

Tetrahedron Lett 44:1507–1509

34. van Buijtenen J, Meuldijk J, Vekemans JAJM, Hulshof LA,

Koojiman H, Spek AL (2006) Dinuclear ruthenium complexes

bearing dicarboxylate and phosphine ligands. acceptorless cat-

alytic dehydrogenation of 1-phenylethanol. Organometallics

25:873–881

35. Morton D, Cole-Hamilton DJ (1987) Rapid thermal hydrogen

production from alcohols catalysed by [Rh(2,20-bipyridyl)2]Cl.J Chem Soc Chem Commun 248–249. doi:

10.1039/C4EE00389F

36. Morton D, Cole-Hamilton DJ (1988) Molecular hydrogen

complexes in catalysis: highly efficient hydrogen production

from alcoholic substrates catalysed by ruthenium complexes.

J Chem Soc Chem Commun 1154–1156. doi:

10.1039/C39880001154

37. Morton D, Cole-Hamilton DJ, Utuk ID, Paneque-Sosa M,

Lopez-Poveda M (1989) Hydrogen production from ethanol

catalysed by group 8 metal complexes. J Chem Soc Dalton

Trans 489–495. doi:10.1039/DT9890000489

38. Delgado-Lieta E, Luke MA, Jones RF, Cole-Hamilton DJ (1982)

The photochemical decomposition of alcohols catalysed by

tri(isopropyl) phosphine complexes of rhodium(I). Polyhedron

1:839–840

39. Arakawa H, Sugi Y (1981) The photocatalytic dehydrogenation

of 2-propanol by using RhCl(PPh3)3. Chem Lett 10:1323–1326

40. Matsubara T, Saito Y (1994) Catalysis of phosphine-coordinated

rhodium(I) complexes for 2-propanol dehydrogenation. J Mol

Catal 92:1–8

41. Adair GRA, Williams JMJ (2005) Oxidant-free oxidation:

ruthenium catalyzed dehydrogenation of alcohols. Tetrahedron

Lett 46:8233–8235

42. Junge H, Beller M (2005) Ruthenium-catalyzed generation of

hydrogen from iso-propanol. Tetrahedron Lett 46:1031–1034

43. Junge H, Loges B, Beller M (2007) Novel improved ruthenium

catalysts for the generation of hydrogen from alcohols. Chem

Commun 522–524. doi:10.1039/DT9890000489

44. Donnelly KF, Segarra C, Shao LX, Suen R, Muller-Bunz H,

Albrecht M (2015) Adaptive N-mesoionic ligands anchored to a

triazolylidene for ruthenium-mediated (de)hydrogenation catal-

ysis. Organometallics 34:4076–4084

45. Valencia M, Muller-Bunz H, Gossage RA, Albrecht M (2016)

Enhanced product selectivity promoted by remote metal coor-

dination in acceptor-free alcohol dehydrogenation catalysis.

Chem Commun 52:3344–3347

46. Delgado-Rebollo M, Canseco-Gonzalez D, Hollering M,

Mueller-Bunz H, Albrecht M (2014) Synthesis and catalytic

alcohol oxidation and ketone transfer hydrogenation activity of

donor-functionalized mesoionic triazolylidene ruthenium(II)

complexes. Dalton Trans 43:4462–4473

47. Sabater S, Muller-Bunz H, Albrecht M (2016) Carboxylate-

functionalized mesoionic carbene precursors: decarboxylation,

ruthenium bonding, and catalytic activity in hydrogen transfer

reactions. Organometallics 35:2256–2266

48. Nordstrøm LU, Vogt H, Madsen R (2008) Amide synthesis from

alcohols and amines by the extrusion of dihydrogen. J Am Chem

Soc 130:17672–17673

49. Dam JH, Osztrovszky G, Nordstrøm LU, Madsen R (2010)

Amide synthesis from alcohols and amines catalyzed by ruthe-

nium N-heterocyclic carbene complexes. Chem Eur J

16:6820–6827

50. Makarov IS, Fristrup P, Madsen R (2012) Mechanistic Investi-

gation of the ruthenium–N-heterocyclic carbene catalyzed ami-

dation of amines with alcohols. Chem Eur J 18:15683–15692

51. Sølvhøj A, Madsen R (2011) Dehydrogenative coupling of

primary alcohols to form esters catalyzed by a ruthenium

N-heterocyclic carbene complex. Organometallics

30:6044–6048

52. Makarov IS, Madsen R (2013) Ruthenium-catalyzed self-cou-

pling of primary and secondary alcohols with the liberation of

dihydrogen. J Org Chem 78:6593–6598

53. Tseng KNT, Kampf JW, Szymczak NK (2013) Base-free

acceptorless and chemoselective alcohol dehydrogenation

P. Pandey et al.

123

Page 17: Acceptorless Alcohol Dehydrogenation: A Mechanistic

catalyzed by an amide-derived N, N, N-ruthenium(II) hydride

complex. Organometallics 32:2046–2049

54. Tseng K-NT, Kampf JW, Szymczak NK (2015) Mechanism of

N, N, N-amide ruthenium(II) hydride mediated acceptorless

alcohol dehydrogenation: inner-sphere b–H elimination versus

outer-sphere bifunctional metal–ligand cooperativity. ACS Catal

5:5468–5485

55. Robert C (2005) The organometallic chemistry of the transition

metals. Wiley, Hoboken, pp 159–180

56. Gladiali S, Alberico E (2004) Transferhydrogenations. In: Beller

M, Bolm C (eds) Transition metals for organic synthesis:

Building Blocks and Fine Chemicals, vol 2, 2nd edn. Wiley,

Weinheim, pp 145–166

57. Yi CS, He Z, Guzei IA (2001) Transfer hydrogenation of car-

bonyl compounds catalyzed by a ruthenium-acetamido complex:

evidence for a stepwise hydrogen transfer mechanism. Orga-

nometallics 20:3641–3643

58. Ritter JCM, Bergman RG (1998) A useful method for preparing

iridium alkoxides and a study of their catalytic decomposition

by iridium cations: a new mode of b-hydride elimination for

coordinatively saturated metal alkoxides. J Am Chem Soc

120:6826–6827

59. Blum O, Milstein D (1995) Mechanism of a directly observed b-hydride elimination process of iridium alkoxo complexes. J Am

Chem Soc 117:4582–4594

60. Bryndza HE, Calabrese JC, Marsi M, Roe DC, Tam W, Bercaw

JE (1986) b-Hydride elimination from methoxo versus ethyl

ligands: thermolysis of (DPPE)Pt(OCH3)2, (DPPE)Pt(CH2-

CH3)(OCH3) and (DPPE)Pt(CH2CH3)2. J Am Chem Soc

108:4805–4813

61. Maggi A, Madsen R (2012) Dehydrogenative synthesis of imi-

nes from alcohols and amines catalyzed by a ruthenium

N-heterocyclic carbene complex. Organometallics 3:451–455

62. Pamies O, Backvall JE (2001) Studies on the mechanism of

metal-catalyzed hydrogen transfer from alcohols to ketones.

Chem Eur J 7:5052–5058

63. Bryndza HE, Tam W (1998) Monomeric metal hydroxides,

alkoxides, and amides of the late transition metals: synthesis,

reactions, and hermochemistry. Chem Rev 88:1163–1188

64. Grotjahn DB, Incarvito CD, Rheingold AL (2001) Combined

effects of metal and ligand capable of accepting a proton or

hydrogen bond catalyze anti-markovnikov hydration of terminal

alkynes. Angew Chem Int Ed 113:4002–4005

65. Erdogan G, Grotjahn DB (2009) Mild and selective deuteration

and isomerization of alkenes by a bifunctional catalyst and

deuterium oxide. J Am Chem Soc 131:10354–10355

66. Grotjahn DB (2005) Bifunctional organometallic catalysts

involving proton transfer or hydrogen bonding. Chem Eur J

11:7146–7153

67. van der Vlugt JI (2012) Cooperative catalysis with first-row late

transition metals. Eur J Inorg Chem 2012:363–375

68. Evans DJ, Pickett CJ (2003) Chemistry and the hydrogenases.

Chem Soc Rev 32:268–275

69. Shibasaki M, Kanai K, Matsunaga S, Kumagai N (2009) Recent

progress in asymmetric bifunctional catalysis using multi-

metallic systems. Acc Chem Res 42:1117–1127

70. Gray TG, Veige AS, Nocera DG (2004) Cooperative bimetallic

reactivity: hydrogen activation in two-electron mixed-valence

compounds. J Am Chem Soc 126:9760–9768

71. Steiman TJ, Uyeda C (2015) Reversible substrate activation and

catalysis at an intact metal–metal bond using a redox-active

supporting ligand. J Am Chem Soc 137:6104–6110

72. Broussard ME, Juma B, Train SG, Peng WJ, Laneman SA,

Stanley GG (1993) A bimetallic hydroformylation catalyst: high

regioselectivity and reactivity through homobimetallic cooper-

ativity. Science 260:1784–1788

73. Ikariya T, Shibasaki M (eds) (2011) Bifunctional molecular

catalysis. Springer, Berlin

74. Gunanathan C, Milstein D (2011) Bond activation by metal–

ligand cooperation: design of ‘‘Green’’ catalytic reactions based

on aromatization–dearomatization of pincer complexes. Top

Organomet Chem 37:55–84

75. Cundari RT, Klinckman TR, Wolczanski PT (2002) Carbon–

hydrogen bond activation by titanium imido complexes. Com-

putational evidence for the role of alkane adducts in selective C–

H activation. J Am Chem Soc 124:1481–1487

76. Gunanathan C, Milstein D (2011) Metal–ligand cooperation by

aromatization–dearomatization: a new paradigm in bond acti-

vation and ‘‘Green’’ catalysis. Acc Chem Res 44:588–602

77. Grutzmacher H (2008) Cooperating ligands in catalysis. Angew

Chem Int Ed 47:1814–1818

78. Himo F, Eriksson LA, Maseras F, Siegbahn PEM (2000) Cat-

alytic mechanism of galactose oxidase: a theoretical study. J Am

Chem Soc 122:8031–8036

79. Parkin G (2004) Synthetic analogues relevant to the structure

and function of zinc enzymes. Chem Rev 104:699–768

80. Matthews BW (1988) Structural basis of the action of ther-

molysin and related zinc peptidases. Acc Chem Res 21:333–340

81. Lipscomb WN, Strater N (1996) Recent advances in zinc

enzymology. Chem Rev 96:2375–2434

82. Christianson DW, Lipscomb WN (1989) Carboxypeptidase A.

Acc Chem Res 22:62–69

83. Zell T, Milstein D (2015) Hydrogenation and dehydrogenation

iron pincer catalysts capable of metal–ligand cooperation by

aromatization/dearomatization. Acc Chem Res 48:1979–1994

84. Mukherjee A, Srimani D, Chakraborty S, Ben-David Y, Milstein D

(2015) Selective hydrogenation of nitriles to primary amines cat-

alyzed by a cobalt pincer complex. J AmChem Soc 137:8888–8891

85. Zhang G, Hanson SK (2013) Cobalt-catalyzed acceptorless

alcohol dehydrogenation: synthesis of imines from alcohols and

amines. Org Lett 15:650–653

86. Ghatak T, Sarkar M, Dinda S, Dutta I, Rahaman SMW, Bera JK

(2012) Olefin oxygenation by water on an iridium center. J Am

Chem Soc 137:6168–6171

87. Daw P, Sinha A, Rahaman SMW, Dinda S, Bera JK (2012)

Bifunctional water activation for catalytic hydration of

organonitriles. Organometallics 31:3790–3797

88. Tolman WB (eds) (2006) Activation of small molecules:

Organometallics and Bioinorganic perspectives. Wiley-VCH:

Weinheim, Germany

89. Yagi M, Kaneko M (2001) Molecular catalysts for water oxi-

dation. Chem Rev 101:21–36

90. Gelman D, Musa S (2012) Coordination versatility of sp3-hy-

bridized pincer ligands toward ligand–metal cooperative catal-

ysis. ACS Catal 2:2456–2466

91. Jimenez MV, Fernandez-Tornos J, Perez-Torrente JJ, Modrego

FJ, Winterle S, Cunchillos C, Lahoz FJ, Oro LA (2011) Irid-

ium(I) complexes with hemilabile N-heterocyclic carbenes:

efficient and versatile transfer hydrogenation catalysts. Orga-

nometallics 30:5493–5508

92. Hetterscheid DGH, van der Vlugt JI, de Bruin B, Reek JNH

(2009) Water splitting by cooperative catalysis. Angew Chem

Int Ed 48:8178–8181

93. Scharf A, Goldberg I, Vigalok A (2013) Evidence for metal–

ligand cooperation in a Pd–PNF pincer-catalyzed cross-cou-

pling. J Am Chem Soc 135:967–970

94. Khusnutdinova JR, Milstein D (2015) Metal–ligand cooperation.

Angew Chem Int Ed 54:12236–12273

95. Haack KJ, Hashiguchi S, Fujii A, Ikariya T, Noyori R (1997)

The catalyst precursor, catalyst, and intermediate in the RuII-

promoted asymmetric hydrogen transfer between alcohols and

ketones. Angew Chem Int Ed 36:285–288

Acceptorless Alcohol Dehydrogenation: A Mechanistic Perspective

123

Page 18: Acceptorless Alcohol Dehydrogenation: A Mechanistic

96. Ohkuma T, Ooka H, Ikariya T, Noyori R (1995) Preferential

hydrogenation of aldehydes and ketones. J Am Chem Soc

117:10417–10418

97. Noyori R, Ohkuma T (2001) Asymmetric catalysis by archi-

tectural and functional molecular engineering: practical chemo-

and stereoselective hydrogenation of ketones. Angew Chem Int

Ed 40:40–73

98. Noyori R (2002) Asymmetric catalysis: science and opportuni-

ties (Nobel Lecture). Angew Chem Int Ed 41:2008–2022

99. Noyori R, Sandoval CA, Muniz K, Ohkuma T (2005) Metal–

ligand bifunctional catalysis for asymmetric hydrogenation. Phil

Trans R Soc A 363:901–912

100. Zweifel T, Naubron JV, Grutzmacher H (2009) Catalyzed

dehydrogenative coupling of primary alcohols with water,

methanol, or amines. Angew Chem Int Ed 48:559–563

101. Zweifel T, Naubron JV, Buttner T, Ott T, Grutzmacher H (2008)

Ethanol as hydrogen donor: highly efficient transfer hydro-

genations with rhodium(I) amides. Angew Chem Int Ed

47:3245–3249

102. Maire P, Bettner T, Breher F, Floch PL, Grutzmacher H (2005)

Heterolytic splitting of hydrogen with rhodium(I) amides.

Angew Chem Int Ed 44:6318–6323

103. Rodrıguez-Lugo RE, Trincado M, Vogt M, Tewes F, Santiso-

Quinones G, Grutzmacher H (2013) A homogeneous transition

metal complex for clean hydrogen production from methanol–

water mixtures. Nat Chem 5:342–347

104. Baratta W, Bossi G, Putignano E, Rigo P (2011) Pincer and

diamine Ru and Os diphosphane complexes as efficient catalysts

for the dehydrogenation of alcohols to ketones. Chem Eur J

17:3474–3481

105. Nielsen M, Junge H, Kammer A, Beller M (2012) Towards a

green process for bulk-scale synthesis of ethyl acetate: efficient

acceptorless dehydrogenation of ethanol. Angew Chem Int Ed

51:5711–5713

106. Alberico E, Sponholz P, Cordes C, Nielsen M, Drexler HJ,

Baumann W, Junge H, Beller M (2013) Selective hydrogen

production from methanol with a defined Iron pincer catalyst

under mild conditions. Angew Chem Int Ed 52:14162–14166

107. Chakraborty S, Lagaditis PO, Forster M, Bielinski EA, Hazari

N, Holthausen MC, Jones WD, Schneider S (2014) Well–de-

fined Iron catalysts for the acceptorless reversible dehydro-

genation–hydrogenation of alcohols and ketones. ACS Catal

4:3994–4003

108. Azerraf C, Gelman D (2009) New shapes of PC(sp3)P pincer

complexes. Organometallic 28:6578–6584

109. Musa S, Shaposhnikov I, Cohen S, Gelman D (2011) Ligand–

metal cooperation in PCP pincer complexes: rational design and

catalytic activity in acceptorless dehydrogenation of alcohols.

Angew Chem Int Ed 50:3533–3537

110. Azerraf C, Shpruhman A, Gelman D (2009) Diels–Alder

cycloaddition as a new approach toward stable PC(sp3)P-met-

alated compounds. Chem Commun 466–468. doi:

10.1039/B815051F

111. Chakraborty S, Piszel PE, Brennessel WW, Jones WD (2015) A

single nickel catalyst for the acceptorless dehydrogenation of

alcohols and hydrogenation of carbonyl compounds. Organo-

metallics 34:5203–5206

112. Blum Y, Shvo Y (1984) Catalytically reactive ruthenium inter-

mediates in the homogeneous oxidation of alcohols to esters. Isr

J Chem 24:144–148

113. Blum Y, Czarkie D, Rahamim Y, Shvo Y (1985) (Cyclopenta-

dienone) ruthenium carbonyl complexes—a new class of homo-

geneous hydrogenation catalysts. Organometallics 4:1459–1461

114. Karvembu R, Prabhakaran R, Natarajan K (2005) Shvo’s

diruthenium complex: a robust catalyst. Coord Chem Rev

249:911–918

115. Shvo Y, Czarkie D, Rahamim Y, Chodosh DF (1986) A new

group of ruthenium complexes: structure and catalysis. J Am

Chem Soc 108:7400–7402

116. Samec JSM, Backvall JE, Andersson PG, Brandt P (2006)

Mechanistic aspects of transition metal-catalyzed hydrogen

transfer reactions. Chem Soc Rev 35:237–248

117. Gunanathan C, Ben-David Y, Milstein D (2007) Direct synthesis

of amides from alcohols and amines with liberation of H2.

Science 317:790–792

118. GunanathanC, ShimonLJW,MilsteinD (2009)Direct conversion

of alcohols to acetals and H2 catalyzed by an acridine-based

ruthenium pincer complex. J Am Chem Soc 131:3146–3147

119. Gnanaprakasam B, Zhang J, Milstein D (2010) Direct synthesis

of imines from alcohols and amines with liberation of H2.

Angew Chem Int Ed 49:1468–1471

120. Gunanathan C, Milstein D (2008) Selective synthesis of primary

amines directly from alcohols and ammonia. Angew Chem Int

Ed 47:8661–8864

121. Ben-Ari E, Leitus G, Shimon LJW, Milstein D (2006) Metal–

ligand cooperation in C–H and H2 activation by an electron-rich

PNP Ir(I) system: facile ligand dearomatization–aromatization

as key steps. J Am Chem Soc 128:15390–15391

122. Balaraman E, Khaskin E, Leitus G, Milstein D (2013) Catalytic

transformation of alcohols to carboxylic acid salts and H2 using

water as the oxygen atom source. Nat Chem 5:122–125

123. Zhang J, Gandelman M, Shimon LJW, Rozenberg H, Milstein D

(2004) Electron rich, bulky ruthenium PNP-type complexes.

Acceptorless catalytic alcohol dehydrogenation. Organome-

tallics 23:4026–4033

124. Zhang J, Balaraman E, Leitus G, Milstein D (2011) Electron-

rich PNP- and PNN-type ruthenium(II) hydrido borohydride

pincer complexes. Synthesis, structure, and catalytic dehydro-

genation of alcohols and hydrogenation of esters. Organome-

tallics 30:5716–5724

125. Fujita KI, Tanino N, Yamaguchi R (2007) Ligand-promoted

dehydrogenation of alcohols catalyzed by Cp*Ir complexes. A

new catalytic system for oxidant-free oxidation of alcohols. Org

Lett 9:109–111

126. Kawahara R, Fujita KI, Yamaguchi R (2012) Cooperative

catalysis by iridium complexes with a bipyridonate ligand:

versatile dehydrogenative oxidation of alcohols and reversible

dehydrogenation–hydrogenation between 2-propanol and ace-

tone. Angew Chem Int Ed 51:12790–12794

127. Kawahara R, Fujita KI, Yamaguchi R (2012) Dehydrogenative

oxidation of alcohols in aqueous media using water-soluble and

reusable Cp*Ir catalysts bearing a functional bipyridine ligand.

J Am Chem Soc 134:3643–3646

128. Yamaguchi R, Ikeda C, Takahashi Y, Fujita KI (2009) Homo-

geneous catalytic system for reversible dehydrogenation–hy-

drogenation reactions of nitrogen heterocycles with reversible

interconversion of catalytic species. J Am Chem Soc

131:8410–8412

129. Fujita KI, Yoshida T, Imori Y, Yamaguchi R (2011) Dehydro-

genative oxidation of primary and secondary alcohols catalyzed

by a Cp*Ir complex having a functional C, N-chelate ligand.

Org Lett 13:2278–2281

130. Zeng G, Sakaki S, Fujita KI, Sano K, Yamaguchi R (2014)

Efficient catalyst for acceptorless alcohol dehydrogenation:

interplay of theoretical and experimental studies. ACS Catal

4:1010–1020

131. Nieto I, Livings MS, Sacci JBI, Reuther LE, Zeller M, Papish

ET (2011) Transfer hydrogenation in water via a ruthenium

catalyst with OH groups near the metal center on a bipy scaf-

fold. Organometallics 30:6339–6342

132. Li H, Wang X, Huang F, Lu G, Jiang J, Wang Z-X (2011)

Computational study on the catalytic role of pincer

P. Pandey et al.

123

Page 19: Acceptorless Alcohol Dehydrogenation: A Mechanistic

ruthenium(II)-PNN complex in directly synthesizing amide from

alcohol and amine: the origin of selectivity of amide over ester

and imine. Organometallics 30:5233–5247

133. Cho D, Ko KC, Lee JY (2013) Catalytic mechanism for the

ruthenium-complex catalyzed synthesis of amides from alcohols

and amines: a DFT study. Organometallics 32:4571–4576

134. Saha B, Sengupta G, Sarbajna A, Dutta I, Bera JK (2014) Amide

synthesis from alcohols and amines catalyzed by a RuII–N-

heterocyclic carbene (NHC)–carbonyl complex. J Organomet

Chem 771:124–130

135. Prokopchuk DE, Morris RH (2012) Inner-sphere activation,

outer-sphere catalysis: theoretical study on the mechanism of

transfer hydrogenation of ketones using iron(II) PNNP enea-

mido complexes. Organometallics 31:7375–7385

136. Wwn O, Morris RH (2013) Ester hydrogenation catalyzed by a

ruthenium(II) complex bearing an N-heterocyclic carbene teth-

ered with an ‘‘NH2’’ group and a DFT study of the proposed

bifunctional mechanism. ACS Catal 3:32–40

137. Mueller JA, Goller CP, Sigman MS (2004) Elucidating the

significance of b-hydride elimination and the dynamic role of

acid/base chemistry in a palladium–catalyzed aerobic oxidation

of alcohols. J Am Chem Soc 126:9724–9734

138. Martınez-Prieto LM, Avila E, Palma P, Alvarez E, Campora J

(2015) b-Hydrogen elimination reactions of nickel and palla-

dium methoxides stabilised by PCP pincer ligands. Chem Eur J

21:9833–9849

139. Casey CP, Johnson JF (2003) Kinetic isotope effect evidence for

a concerted hydrogen transfer mechanism in transfer hydro-

genations catalyzed by [p-(Me2CH)C6H4Me]Ru-

(NHCHPhCHPhNSO2C6H4-p-CH3). J Org Chem 68:1998–2001

140. Hou C, Zhang Z, Zhao C, Ke Z (2016) DFT study of accep-

torless alcohol dehydrogenation mediated by ruthenium pincer

complexes: ligand tautomerization governing metal ligand

cooperation. Inorg Chem. doi:10.1021/acs.inorgchem.6b00723

141. Dutta I, Sengupta G, Bera JK (2015) Reactivity and catalysis at

sites trans to the [Ru–Ru] bond. Top Organomet Chem. doi:

10.1007/3418_2015_162

142. Sean CE, Hadzovic A, Morris RH (2004) Mechanisms of the

H2-hydrogenation and transfer hydrogenation of polar bonds

catalyzed by ruthenium hydride complexes. Coord Chem Rev

248:2201–2237

143. Dutta I, Sarbajna A, Pandey P, Rahaman SMW, Singh K, Bera

JK (2016) Acceptorless dehydrogenation of alcohols on a

diruthenium(II, II) platform. Organometallics 35:1505–1513

144. Zhang J, Leitus G, Ben-David Y, Milstein D (2005) Facile

conversion of alcohols into esters and dihydrogen catalyzed by

new ruthenium complexes. J Am Chem Soc 127:10840–10841

145. Spasyuk D, Gusev DG (2012) Acceptorless dehydrogenative

coupling of ethanol and hydrogenation of esters and imines.

Organometallics 31:5239–5242

146. Friedrich A, Schneider S (2009) Acceptorless dehydrogenation

of alcohols: perspectives for synthesis and H2 storage. Chem-

CatChem 1:72–73

147. Murahashi SI, Naota T, Hirai N (1993) Aerobic oxidation of

alcohols with ruthenium–cobalt bimetallic catalyst in the pres-

ence of aldehydes. J Org Chem 58:7318–7319

148. Saha B, Rahaman SMW, Daw P, Sengupta G, Prof Bera JK

(2014) Metal–ligand cooperation on a diruthenium platform:

selective imine formation through acceptorless dehydrogenative

coupling of alcohols with amines. Chem Eur J 20:6542–6551

149. Patra SK, Bera JK (2007) C–C bond forming reaction through

aldol-type addition mediated by a [Ru2(CO)4]2? core. Organo-

metallics 26:2598–2603

150. Stevens CV, Vekemans W, Moonen K, Rammeloo T (2003)

Synthesis of 4-phosphono-b-lactams via phosphite addition to

acyliminium salts. Tetrahedron Lett 44:1619–1622

Acceptorless Alcohol Dehydrogenation: A Mechanistic Perspective

123