You are on page 1of 15

pubs.acs.

org/acscatalysis Review

Recent Advances in Liquid Organic Hydrogen Carriers: An Alcohol-


Based Hydrogen Economy
Vinita Yadav, Ganesan Sivakumar, Virendrakumar Gupta, and Ekambaram Balaraman*
Cite This: ACS Catal. 2021, 11, 14712−14726 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Energy storage and the use of abundantly available


feedstock without contributing to the carbon footprint are two
significant global challenges. In this regard, the development of
Downloaded via UNIV OF ABERDEEN on March 27, 2022 at 16:30:36 (UTC).

high-performance, low-cost, sustainable, and environmentally


friendly energy storage and production systems is crucial to fulfill
the growing energy demands of the current society. The use of
hydrogen will diversify energy sources as it significantly reduces
greenhouse gas emissions and environmental pollution during
energy conversion. Although the hydrogen economy is quite
beneficial, hydrogen storage is still very challenging, and the
existing methods suffer from a lot of problems and drawbacks. The
conventional liquefaction and compression hydrogen storage
technologies are associated with several challenges, including low storage density, boil-off losses, relatively high costs, and safety
and transportation concerns. In recent years, liquid organic hydrogen carrier (LOHC) systems have attained a lot of importance as a
substitute for the traditional storage methods. Hydrogen storage and transport using LOHCs are based on two-step cycles, such as
(i) loading/storage of hydrogen by catalytic hydrogenation of H2-lean compounds and (ii) unloading/releasing hydrogen by
dehydrogenating the resulting H2-rich liquids. Since alcohols are widely accessible from various industrial processes or even from
biomass-derived precursors, the catalytic acceptorless dehydrogenation of alcohols is an attractive approach for future hydrogen
storage applications. Hence, the catalytic dehydrogenation-hydrogenation of alcohols can be used for the development of alcohol-
based LOHC systems which are economical, safe, and easy to handle. Further, they are similar to crude oils under ambient
conditions and thus are suitable for use in the current energy infrastructure. This Review covers several essential aspects of these
developing efficient and abundantly available LOHC systems for efficient hydrogen storage and transport applications. Additionally,
reversible LOHC systems based on the catalytic dehydrogenation-hydrogenation of alcohols and their corresponding carbonyl
compounds have been discussed.
KEYWORDS: alcohol, acceptorless dehydrogenation, dehydrogenative coupling, homogeneous catalysis, hydrogen, hydrogenation, LOHCs

■ INTRODUCTION
The rapid exhaustion of fossil fuels, anthropogenic emissions,
decades ago. In 1971, Jones initially proposed that liquid
hydrogen could replace fossil fuels in the 21st century.6 Later,
and ever-increasing energy consumption has led to an Salzano, Momirlan, and Bockris independently described a
increasing interest in the future energy system based on hydrogen fuel economy and hydrogen storage concepts.7−9
renewable resources. Indeed, renewable energy sources, such as Despite hydrogen’s potential as a clean and efficient energy
wind, solar, geothermal, and biomass have attracted consid- carrier, the production, storage, delivery, and end-use of the gas
erable attention as alternatives to fossil fuels. However, some of pose significant challenges. Hydrogen is primarily produced
these sources, namely, wind and solar, are naturally from fossil fuels such as natural gas and coal. However, in
intermittent and fluctuating. Hence, it is necessary to find recent years, renewable and sustainable energy sources have
alternative energy sources that can fulfill our future energy shown considerable potential in hydrogen production. For
demands. In this context, hydrogen is considered a clean and example, photocatalytic, thermochemical, or electrolytic water
efficient energy carrier because of its high energy content (120
MJ/kg) and clean-burning properties compared with fossil Received: July 21, 2021
fuels (Table 1).1−3 As an energy carrier, hydrogen can be used Revised: November 11, 2021
for various fuel cell applications, such as mobile, stationary, and Published: November 23, 2021
portable power applications, without any greenhouse gas
emissions.4,5 Utilization of hydrogen as an energy source and
carrier is not a new concept and was recommended several

© 2021 American Chemical Society https://doi.org/10.1021/acscatal.1c03283


14712 ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Table 1. A Comparison of the Energy Contents of Different government.27,28 Moreover, the H2 released from these
Fuels1 compounds cannot be efficiently reloaded since the liquid
carrier gets consumed. However, in recent decades, incredible
sample no. fuel energy content (MJ/kg)
findings and significant milestones have been achieved in the
1 molecular hydrogen 120 catalytic hydrogenation of CO2 to methanol both under
2 liquefied natural gas 54.4 homogeneous and heterogeneous catalytic conditions (see
3 propane 49.6 Supporting Information).29−32 Indeed, capturing the CO2 and
4 aviation gasoline 46.8 converting it to methanol can reduce the carbon footprint and
5 automotive gasoline 46.4 environmental feasibility of utilization.
6 automotive diesel 45.6 An alternative hydrogen carrier system that has gained much
7 ethanol 29.6 attention is the liquid organic hydrogen carrier (LOHC). By
8 methanol 19.7 definition, LOHCs are liquids or low-melting solids which
9 coke 27 undergo reversible hydrogenation and dehydrogenation in the
10 wood (dry) 16.2 presence of a suitable catalytic system. LOHCs are compatible
with present transport and refueling infrastructures and permit
splitting into hydrogen (H2) and oxygen (O2), and electricity high gravimetric and volumetric hydrogen storage under
from wind, solar, geothermal, or biomass are also currently ambient conditions. The LOHC system consists of a pair of
used to produce hydrogen.10−13 hydrogen-lean and hydrogen-rich organic compounds. Hydro-
Despite extensive progress toward hydrogen production, the gen is stored by converting H2-lean organic compounds to H2-
storage and transportation of hydrogen have several rich organic compounds in the catalytic hydrogenation
limitations, including a low volumetric energy density reaction, whereas the reverse dehydrogenation reaction of
(0.0108 MJL−1); safety issues; and the cost of cryogenic, H2-rich systems results in hydrogen release.33,34 This overall
high-pressure compression cylinders or tanks.14−18 It is process is reversible, and only hydrogen is released. The
therefore highly desirable to develop new methods for LOHC concept is schematically illustrated in Scheme 2.
convenient transportation and storage of hydrogen. Over the Although several hydrogenation and dehydrogenation reac-
past decades, many types of hydrogen storage technologies tions of organic compounds are reported in the literature,35−37
have been investigated. For instance, hydrogen adsorption as there are certain requirements for these organic compounds to
metal hydrides, in metal−organic frameworks (MOFs), and in become efficient LOHC systems:
nanostructured materials have been studied.19−22 However,
these methods are limited by low hydrogen storage capacities High hydrogen storage capacity (volumetric and
(HSCs), harsh conditions, high cost, and low energy efficiency. gravimetric storage capacities >56 kg/m3 and >6 wt %,
The storage of hydrogen in chemical bonds of tiny organic respectively).
molecules, particularly organic liquids, is an appealing avenue Low production cost and unlimited availability for a low
for future hydrogen carrier systems. Methanol,23 form- price.
aldehyde,24 and formic acid25 based hydrogen carrier systems Compatibility with existing energy infrastructure.
have been gaining considerable attention in the past few years. All compounds should have low melting point (<−30
The simplest alcohol, methanol, is emerging as a potential °C) so that the use of external solvents can be avoided.
candidate for an organic hydrogen carrier because of its 12.6 wt The high boiling point (>300 °C) of the LOHC system
% hydrogen content, ease of handling, and convenient is necessary to simplify hydrogen purification by
production of hydrogen .26 The dehydrogenation of aqueous condensation.
methanol is a three-step process. Initially, formaldehyde is Should be durable to undergo selective hydrogenation
formed via dehydrogenation of methanol; then hydration of and dehydrogenation for more cycles.
formaldehyde yields methanediol, which again undergoes The LOHC compounds should be safe and nontoxic,
dehydrogenation to afford formic acid; and the final step with high toxicological and eco-toxicological profiles
results in CO2 production from formic acid (Scheme 1). This during transportation and usage.
Several promising LOHC systems based on hydrocarbons
Scheme 1. Description of the Basic Steps Involved in the
and N-heterocycles, such as decalin-naphthalene, methylcyclo-
Dehydrogenation of Aqueous Methanol
hexane-toluene, N-ethylcarbazole and perhydro-N-ethylcarba-
zole, 2-methyl-1,2,3,4-tetrahydroquinoline, and 2,6-dimethyl-
decahydro-1,5-naphthyridine have been reported.4,17,37−40
However, most systems have limitations, including high
hydrogen pressure and temperatures requirements for hydro-
genation and dehydrogenation reactions, low HSCs, and high
LOHC costs.
overall process releases three molecules of hydrogen and CO2. Since alcohols are easily manageable, inexpensive, and
Similarly, formaldehyde and formic acid also release CO2 by widely available from various industrial processes and even
dehydrogenation. Although these hydrogen carrier systems are from renewable resources like lignocellulose biomass, they are
quite promising, they suffer from several drawbacks, including a good candidate for hydrogen storage applications.41,42
the release of CO2 from methanol and formaldehyde, toxicity, Alcohols undergo catalytic dehydrogenation to produce
and the low hydrogen storage capacity of formic acid (4.4 wt carbonyl derivatives as hydrogen-lean byproducts, which can
%). In the context of environment viability, the hydrogen be converted into the initial alcohols by hydrogenation.
carrier systems should have an HSC of 5.0 and 6.5 wt %, Further, the dehydrogenation of alcohols can be performed
respectively, set by the European Union and the U.S. at lower temperatures as compared with N-heterocycles.
14713 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Scheme 2. Schematic Representation of Hydrogen Storage Concept Using LOHC Technology

Hence, alcohols are exciting candidates for sustainable As discussed earlier, methanol has a high hydrogen storage
hydrogen production that can also act as rechargeable liquid capacity and can be a good candidate for hydrogen storage
organic hydrogen carriers (LOHCs). This Review will discuss applications. The dehydrogenation of aqueous methanol
transition-metal catalyzed acceptorless dehydrogenation of results in gaseous product (CO2), whereas anhydrous
bioalcohols from the perspective of hydrogen storage in methanol leads to formaldehyde and methyl formate (MF).
alcohols. Additionally, reversible LOHC systems have been In the case of water contamination, formic acid and CO2 are
discussed with various catalytic dehydrogenation-hydrogena- also produced as minor products. In 1985, the Maitlis group
tion of alcohols and their corresponding carbonyl compounds. reported the dehydrogenative transformation of methanol to


methyl formate by using 0.03 mol % of [Ru(PPh3)3Cl2]
ACCEPTORLESS DEHYDROGENATION OF catalyst at 150 °C for 18 h.43 Mechanistic studies on methanol
ALCOHOLS dehydrogenation with these types of catalysts indicated that
the presence of bulky and poor-electron-donating ligands could
The dehydrogenation of alcohols has been a widely explored accelerate the reaction rate.44 Further, the research group of
catalytic process from the beginning of the 21st century. Both Saito used [Ru(OAc)Cl(PEtPh2)3] as a catalyst along with
noble- and base-metal-based various homogeneous catalysts acetic acid as an additive at 66 °C for this transformation and a
have been investigated to efficiently dehydrogenate alcohols to TON of 34 was obtained after 90 days.45 On the basis of
aldehydes, ketones, carboxylic acids, or esters. In acceptorless several experimental studies and the mechanism postulated by
alcohol dehydrogenation (AAD), alcoholic precursor under- Dobson and Robinson for the dehydrogenation of alcohols,46
goes dehydrogenation, which results in the evolution of Saito and co-workers proposed the mechanism for the
hydrogen molecule is quite interesting. Since reverse hydro- formation of formaldehyde from methanol by dehydrogen-
genation of these aldehydes, esters and carboxylic acids are also ation47 (Scheme 4). In 1988, Cole-Hamilton’s research group
widely explored using the same or different catalytic systems, reported the dehydrogenation of alcohol using [RuH2(N2)-
reversible acceptorless dehydrogenation/hydrogenation reac- (PPh3)3] complex with the help of a base under light
tions could be utilized as promising hydrogen storage systems. irradiation at 150 °C.48,49 In addition to alcohols, diols,
Biomass-derived alcohols (methanol, ethanol, and polyols) are polyols, methanol, and ethanol were also successfully
potential candidates for hydrogen storage applications due to dehydrogenated with a turnover of 37 and 210 h−1,
their relatively high hydrogen content (Scheme 3). Moreover, respectively. The reaction mechanism proceeds via the
liquid alcohols can be utilized directly as a fuel by the existing formation of ruthenium anionic species followed by β-hydride
transportation infrastructure. elimination led to trihydrido species and the corresponding
carbonyl compound. Further, the hydrogen abstraction from
Scheme 3. Alcohol Mainly Used in Hydrogen Storage another alcohol molecule results in the formation of
Applications [RuH2(H2)(PPh3)]. Finally, the catalytic loop completes
with the evolution of molecular hydrogen (Scheme 5). In
1990, another example of methanol dehydrogenation to methyl
acetate was reported by Shinoda and Yamakawa.50
They used a bimetallic Ru(II)−Sn(II) complex [Ru-
(SnCl3)5(PPh3)]3− and MeOH/MeNO2 (1:1) solvent mixture
under inert atmosphere at 65 °C, which showed low catalytic
activity and a turnover number of only 15.7 was obtained after
100 h. Mechanistic studies revealed that the rate-determining
14714 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Scheme 4. Proposed Mechanism for Methanol co-workers developed another molecularly defined catalytic
Dehydrogenation46 system [CpRu(PPh3)2(SnF3)] based on bimetallic Ru/Sn site.
The complex [CpRu(PPh3)2(SnF3)] showed TONs up to 120
under similar reaction conditions (MeOH/MeNO2 (1:1)
solvent mixture at 140 °C). The improved activity might be
attributed to the fluoro-containing ligands compared to
chloride and bromide analogs as fluorine is more electro-
negative, which increases positive charge on the ruthenium
atom.52 Further, in the rate-determining step (β-hydrogen
abstraction), heterolysis of C−H bond might be facilitated by
the formation of intramolecular hydrogen bond, could also
contribute to the increased activity in the fluoro derivatives.
There is much potential research in aqueous-phase methanol-
reforming reactions that have been studied in recent years (see
Table 2, the catalyst for dehydrogenation of methanol). Since
many reviews have focused on the aqueous-phase reforming of
methanol,23,53 such topics are not included.
Ethanol is another promising candidate for use as a LOHCs
due to its high hydrogen output, abundant, economical, and
easy accessibility from renewables. Previously, the generation
of hydrogen from ethanol had been studied.47,48 Since ethanol
is, apart from methanol, an extremely challenging substrates
because of its relatively high energy barrier, a stable catalyst
Scheme 5. Proposed Mechanism for the Dehydrogenation system is required for its dehydrogenation. Further, the
of Alcohols Using [RuH2(N2)(PPh3)3] Complex48 dehydrogenation of ethanol can result in other side reactions,
lowering the reaction efficiency.54 The selection of a suitable
reaction condition is essential in the dehydrogenation of
ethanol. Indeed, ethanol dehydrogenation can lead to three
different products such as acetaldehyde (via AD), ethyl acetate
(via ADC), and acetic acid (Scheme 6).

Scheme 6. Ethanol Dehydrogenation Routes

In 1977, Dobson and Robinson used the ruthenium and


osmium complexes [M(OCORF)2(CO)(PPh3)2] (M= Ru, Os,
RF = CF3, C2F5, or C6F5) for the synthesis of various aldehydes
and ketones from primary and secondary alcohols respectively
via the dehydrogenation reaction.47 The dehydrogenation
step is the dehydrogenation of methanol to formaldehyde. reaction was carried at reflux temperatures in the presence of
Further, it was presumed that the Ru(II)−Sn(II) bimetallic site 12 equiv of trifluoroacetic acid to enhance the reaction. For the
catalyzes the formation of acetic acid by isomerization of primary alcohols, an initial rate of dihydrogen elimination
methyl formate through a four-center interaction of soft Ru(II) increases with the boiling point and ranges from 0.0075 mol
and hard Sn(II) with soft CO group and hard -OMe group s−1 (mol of catalyst) −1 for ethanol to 2.27 mol s−1 (mol of
of methyl formate, respectively. Finally, this leads to migration catalyst)−1 for benzyl alcohol. The catalytic cycle proposed
of the methyl group (from O to C).51 In 2000, Gusevskaya and involves solvolysis of the catalyst and generates the

Table 2. Selected Catalytic Systems Used in the Dehydrogenation of Anhydrous Methanol

catalytic system solvent additive temp (°C) reaction time (h) TON TOF (h−1)
[Ru(PPh3)3Cl2] MeOH − 150 18 65 3.6
[Ru(OAc)Cl(PEtPh2)3] MeOH AcOH 66 90 34 0.4
[RuH2(N2)(PPh3)3] MeOH NaOH 150 2 74 37
[Ru(SnCl3)5(PPh3)]3− MeOH/MeNO2 − 65 100 15.7 0.15
[CpRu(PPh3)2(SnF3)] MeOH/MeNO2 − 140 40 120 3

14715 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

corresponding M-alkoxide species [M(OAlk)(OCORF)(CO)- Scheme 8. Ru-PNP Complexes for Efficient


(PPh3)2] followed by β-hydride elimination to generate the Dehydrogenation of Ethanol to Ethyl Acetate54,60,61
respective aldehyde (or ketone) and form the metal hydride
complex [MH(OCORF)(CO)(PPh3)2]. Further, the trifluoro-
acetic acid reacts with the hydride complex to regenerate the
catalysts and liberate dihydrogen. Later, Rybak and Ziółkow-
ski55 and Jung and Garrou56 independently tried to regenerate
Dobson and Robinson experimental results. However, short-
range alcohols like ethanol, isopropanol, and n-butanol
achieved lower conversions under comparable reaction
conditions using the ruthenium acetate complex. Jung and
Garrou explained that the poor activity and short catalyst
lifespan are due to its thermal decomposition and loss of
trifluoroacetic acid under the catalytic reaction conditions.
Hulshof and co-workers also reported the catalytic acceptorless
dehydrogenation of alcohols into aldehydes and ketones using
the same Robinson catalyst.57 Using 1 mol % catalyst and 12−
18 equiv of trifluoroacetic acid (with respect to catalyst),
various primary and secondary alcohols were dehydrogenated performed under an ordinary atmosphere. Moreover, Beller co-
selectivity to the respective carbonyl compounds in a short workers have reported another method for effective hydrogen
reaction period at 130 °C. However, compared with Dobson generation from aqueous ethanol and bioethanol catalyzed by a
and Robinson’s work, conversion and selectivity obtained from Ru(II)-complex, [Ru(H)(Cl)(CO)(iPr2PEtN(H)EtPiPr2)]
the dehydrogenation of primary alcohols were not appreciated. (3).54 Using this aliphatic Ru-pincer catalyst in the presence
This is attributed to other potential side reactions like aldol of 8 M NaOH at 90 °C, a TON of 1770 after 1 h was achieved.
condensation and decarbonylation of the aldehyde intermedi- Aqueous ethanol dehydrogenation proceeded with good
ate. Despite these encouraging earlier results, considerable selectivity and provided up to 70% acetic acid. Moreover, a
advances in simplified catalyst systems, stabilities, and activities TON of 80 000 was obtained after 4 days, indicating the
were still required for efficient hydrogen production. In 2007, robustness of this catalyst system (Scheme 8).
Beller’s research group reported the dehydrogenation of In 2016, the Milstein group reported another method for
isopropyl alcohol using Ru(II) catalytic system.58 Using C2-alcohols like amino alcohols, acceptorless dehydrogenation
[RuCl2(p-cymene)]2 catalysts along with tetramethylethylene- to their corresponding C2-carboxylic acid salts, which was
diamine (TMEDA), excellent catalyst stability (>250 h) and successfully applied to synthesize a variety of amino acid
high turnover frequencies (TOF) up to 519 h−1 were achieved salts.62 Using 0.1 mol % ruthenium complex 4 in a basic
for the AD of isopropyl alcohol. Further, the same catalyst aqueous medium at 125 °C, a range of amino alcohols were
system was used for the dehydrogenation of ethanol, obtaining shown to have high reactivity and were obtained in good yields
a TOF of 7.6 h−1 after 2 h. Subsequently, the Beller group also (Scheme 9).
reported a combination of [RuH2(PPh3)3CO]/PNPiPr catalytic
system for more efficient hydrogen production from ethanol Scheme 9. Ru(II)-PNN Catalyzed Synthesis of Amino Acid
and isopropyl alcohol under neutral conditions (Scheme 7).59 Salts from Amino Alcohol62

Scheme 7. Ru-PNP Complex for the Efficient


Dehydrogenation of Alcohols to Aldehydes or Ketones59

Interestingly, water is playing the dual role of acting as a


solvent and a nucleophile, from which the oxygen atom of the
carboxylic acid group was derived, with the liberation of
The in situ generated ruthenium catalytic system showed dihydrogen. Besides the precious metal-based catalytic systems,
excellent reactivity in the dehydrogenation of isopropanol and the earth-abundant 3d-metal complexes are also known for the
ethanol under mild conditions with a TOF (after 2 h at 100 acceptorless dehydrogenation of alcohols.63 The research
°C) of 8382 and 1483 h−1 respectively. In 2012, Gusev60 and groups of Jones and Schneider reported the first iron-based
Beller61 groups independently reported the efficient hydrogen acceptorless dehydrogenation of alcohols.64 Using a minimal
generation from ethanol and converted it to ethyl acetate amount of catalyst (0.1−1 mol %) of 5, a range of alcohols
through the ADC (Scheme 8). Under very similar conditions, including aliphatic, primary, and secondary alcohols and diols
both the ruthenium catalysts 1 and 2 efficiently catalyze the were dehydrogenated to their corresponding carbonyl
dehydrogenative coupling of ethanol with excellent TONs derivatives such as aldehydes, ketones, esters, or lactones,
around 15 000 and 17 000, respectively. Also, the product ethyl respectively, in average to good isolated yields (Scheme 10,
acetate was obtained in a close range (77% and 83% yields, Table 3). Further, the catalyst showed good chemoselectivity
respectively). Catalyst 2 is air-stable, and the reaction was in the presence of both primary and secondary diols. Notably,
14716 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Scheme 10. Iron-PNP (5) Complex Catalyzed Acceptorless Dehydrogenation of Alcohols64

Table 3. Overview of Selected Catalytic Systems Used in Ethanol or Polyols Dehydrogenation

the secondary alcohol only underwent dehydrogenation conversion of AE, forming GA in 48% yield with linear
without affecting the primary alcohol. According to the DFT peptide. Further, the improvement in the reaction conditions
studies and control experiments, the active catalyst in the AD/ using catalyst 6 and the other bipyridine-based pincer
ADC reactions was proposed to be a pentacoordinate iron complexes 4 and 8 is shown in the lower conversion of AE.
amido hydride species 5′. By replacing the NEt2 group with an N(H)(tBu) amine group


in the pincer complex 6, the catalytic activity was enhanced.
ALCOHOL−AMINE-BASED LOHC SYSTEMS Thus, using catalyst 9 in the presence of 1.2 mol % KOtBu (2.4
In 2015, Milstein and co-workers reported the ruthenium- equiv relative to 9) and keeping the remaining conditions
catalyzed reversible LOHC system based on amine-alcohols. same, the reaction resulted in 85% conversion of AE and 60%
Acceptorless dehydrogenative amide formation from amino yield of GA with 37 mL (77% yield) hydrogen (using reaction
alcohol (AA) and the reverse hydrogenation using the same 1b). Later, the reverse hydrogenation reaction of GA was
catalyst regenerate AA.65 They used PNN ruthenium pincer performed using complexes 6 and 9. It was found that complex
complexes 4 and 6−9 that efficiently catalyze several 9 was catalytically much more active compared with 6, and
dehydrogenative coupling reactions of alcohols and amines almost quantitatively, conversion of expected product AE was
(Scheme 11). In this study, 2-aminoethanol (AE) was used as obtained with 0.5 mol % of 9 and H2 pressure (50 bar) at 110
an alcohol and an amine, which upon acceptorless dehydrogen- °C. Pleasingly, the mixed products GA and LP obtained via the
ative coupling process can either form piperazine-2,5-dione dehydrogenation reaction were also hydrogenated under a
(glycine anhydride, GA) or the linear oligopeptide (LP). The similar set of conditions, leading to an 85 wt % yield of AE,
formation of GA as a major product is more advantageous which implies that the formation of LP does not impose a
because it has a high hydrogen storage potential of 6.56 wt % major problem in the regeneration of AE.
as compared with LP, which lowers the hydrogen storage The repeated dehydrogenation-hydrogenation cycles were
capacity of AE (for n = 6, 5.46 wt %). The reaction of 1 mmol also carried out without the addition of additional catalysts,
AE in low catalyst loading like 0.5 mol % of catalyst 6 and and 81% of AE was detected even after the third cycle,
KOtBu (1.2 equiv relative to 6) with 4 mL of dioxane solvent indicating that this system has the potential to be used in
at 135 °C for 12 h under argon atmosphere resulted in 78% hydrogen storage (Scheme 12).
14717 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Scheme 11. Reactions of Amino Alcohol-Based Reversible Hydrogen Carrier System65

Scheme 12. Repetitive Dehydrogenation-Hydrogenation of ED and ethanol using 0.2 mol % catalyst 4 and 1 mL of
Cycles dioxane solvent at 135 °C for 24 h gave the expected product
N,N′-diacetylethylenediamine (DAE) in 92% yield with full
conversions of ED and ethanol and the monoamide N-(2-
aminoethyl)-acetamide (AEA) in only 8% yield, as well as H2
in 95% yield (Scheme 13). Further, the reverse hydrogenation
reaction using 0.2 mol % 4 and KOtBu (2.4 equiv relative to 4)
under 70 bar hydrogen pressure at 115 °C for 48 h afforded
In subsequent work, Milstein and co-workers developed a complete conversion of DAE and ED in 92% of yield. Using
new LOHC system using commercially available, abundant,
and inexpensive ethylenediamine (ED) and ethanol as 0.4 mol % catalyst 4 and less dioxane solvent (2 mL), large-
hydrogen carriers.66 In this reaction, bipyridine-based pincer scale hydrogenation of DAE under the same set of conditions
complex 4 showed excellent activity with a catalytic amount of for 10 h resulted in the quantitative amount of ED.
base. For both hydrogenation and dehydrogenation, complex 4 Plausible organic intermediates are shown for the dehydro-
showed excellent conversions. The dehydrogenative coupling genative coupling of ED and ethanol in Scheme 14.

Scheme 13. Ru-PNN Complex (4) Catalyzed Dehydrogenative Coupling of Ethylenediamine with Ethanol66

14718 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Scheme 14. Proposed Pathway for the Dehydrogenative reaction conditions, the ruthenium complex 10, which is the
Coupling of Ethanol and Ethylenediamine (ED) PPh2 analogue of complex 9, was used for the dehydrogenative
coupling of ED and 1,4-butanediol, resulting in 70% yield of
bis-cyclic imide and 84% hydrogen yield. The reverse
hydrogenation of bis-cyclic imide at 40 bar hydrogen pressure
using complex 9 (1 mol %) and KOtBu (3 mol %) in dioxane
resulted in 99% conversion of starting imide and more than
90% yield of 1,4-butanediol and ethylenediamine. A similar
reaction using complex 10 showed conversion of the bis-cyclic
imide in 65% only. The repetitive hydrogenation/dehydrogen-
ation between the spent fuel (bis-cyclic imide) and the charged
fuel (ED/1,4-butanediol mixture) using complexes 9 and 10
showed moderate efficiency for the imide formation up to two
cycles. The ruthenium complex 9 was also used in hydro-
genation of various cyclic imides to form diols and amines in
good to excellent yields.
In seminal work, the research group of Olah and Prakash
described methanol/amine based a reversible hydrogen storage
system. Indeed, hydrogen is generated in the presence of
ruthenium catalyst 3 or 11 via the dehydrogenative coupling of
The research group of Milstein developed a new hydrogen methanol and amine and the reverse hydrogenation reaction
storage system based on the formation of bis-cyclic imide via regenerates the starting CH3OH and amine by the same
dehydrogenative coupling of inexpensive and readily available catalyst, thus closing the cycle.68 Methanol is particularly
starting materials like 1,4-butanediol and ethylenediamine suitable in this type of hydrogen storage system because it
(ED) and then hydrogenation of bis-cyclic imide.67 Using contains 12.6 wt % hydrogen; it is easy to handle; and it can be
ruthenium complex 9 (1 mol %) and KOtBu (2 mol %), the manufactured by a variety of industrial processes. The reaction
dehydrogenative coupling of ED and 1,4-butanediol in dioxane of benzylamine and methanol with catalyst 11 in toluene
(2 mL) at 120 °C for 24 h, given the complete conversion of solvent at 140 °C for 24 h in a closed reactor resulted in an
1,4-butanediol. However, bis-cyclic imide was formed in 55% 88% yield of N,N′-dibenzylurea. While the reverse hydro-
yield and 20% yield of lactone along with other products such genation reaction (60 bar of hydrogen pressure) happened
as oligoamides (Scheme 15). Furthermore, under similar with a complete conversion of N,N′-dibenzylurea to methanol

Scheme 15. Reversible Transformation of 1,4-Butanediol and Ethylenediamine67

14719 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Scheme 16. Acceptorless Dehydrogenative Coupling Affording Formamide Using Methanol68

Scheme 17. Manganese-Catalyzed Acceptorless Dehydrogenative Coupling Affording Formamide Using Methanol69

and benzylamine, this system has only 2.4 wt % H2 storage transition metals for LOHC systems. The dehydrogenative
potential because of the high carbon content of benzylamine. condensation of methanol and N,N′-dimethylethylenediamine
In view of the hydrogen storage capacity, N,N′-dimethylethy- (DED) with 2 mol % PhPNP-complex 12 and KOtBu (4 mol
lenediamine (DED, 6.6 wt %) was the most appropriate amine, %) in 1,4-dioxane at 165 °C for 16 h gave N,N′-(ethane-1,2-
and the reaction of 1 mmol of DED and methanol (4 mmol) diyl)bis(N-methylformamide) in 86% of yield and along with
with 1 mol % 11 and 5 mol % K3PO4 resulted in 86% H2 yield 92% (95.5% purity) yield of H2 (Scheme 17). Assuming that
at 120 °C in toluene solvent for 24 h (Scheme 16). When Ru- the H2 produced during the reaction in a closed system would
MACHO catalyst 3 was used under a similar set of conditions, inhibit dehydrogenation, the evolved hydrogen was released
90% H2 yield was observed along with 95% yield of DED in after 2 h of reaction, and the reaction was then carried out for
the reverse hydrogenation reaction at 60 bar hydrogen another 6 h at 165 °C, which resulted in a higher yield of both
pressure. Moreover, scaling up the reaction (5 mmol) under H2 and the desired product (98% and 97% yield, respectively).
neat conditions (in the absence of solvent) also resulted in In comparison to the ruthenium catalyst 3, this PhPNP-
moderate to good yield of dehydrogenated and hydrogenated complex 12 showed 97% selectivity for a formation of N,N′-
products (76% and 60%, respectively). (ethane-1,2-diyl)bis(N-methylformamide) by the dehydrogen-
More recently, Liu et al.65 demonstrated the utility of an ative condensation of methanol and N,N′-dimethylethylenedi-
earth-abundant manganese pincer complex in developing a amine (DED). Further, the hydrogenation of N,N′-(ethane-
similar hydrogen carrier system as described by Olah and 1,2-diyl)bis(N-methylformamide) required harsh reaction
Prakash. Notably, this is the first report of the catalytic system conditions (180 °C, 80 bar of H2) to obtain N,N′-
based on cost-effective, sustainable, and efficient first-row- dimethylethylenediamine (DED) in 94% yield by catalyst 12,
14720 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Scheme 18. Hydrogenation/Dehydrogenation Reactions Involving Methanol, Ethylene Diamine, and Urea70

Scheme 19. Mechanism of the ADC of Ethylenediamine and Methanol

the same catalyst used in the dehydrogenation reaction. While with ruthenium catalysts 4 and 6, the reaction resulted in 100%
catalyst 13 was found to be the most active for this conversion and 100% yield of ED and methanol after 1.5 and 5
hydrogenation reaction and resulted in 99% yield of DED days. The catalyst 4 showed excellent activity for EU
with catalyst 13 (2 mol %) and KOtBu (2.5 mol %) in dioxane hydrogenation even when the reaction time was reduced to
solvent at 110 °C for 16 h. Milstein and co-workers70 1 d, yielding 81% of ED and 77% methanol. Next, the
developed a similar hydrogen carrier system based on the dehydrogenative coupling of ethylenediamine and methanol
formation of ethylenediamine (ED) and methanol by (1:3) was carried out in a closed 25 mL Young-type Schlenk
unprecedented hydrogenation of ethylene urea (EU) and its tube using 1 mol % 9 and 2.2 mol % KOtBu in dioxane (2 mL)
reverse dehydrogenative coupling reaction in the presence of at 150 °C for 48 h. This reaction yielded 99% H2 and 76% EU,
Ru-catalysts 4, 6, and 9. Earlier, the same group established a as well as intermediates N-(2-aminoethyl)formamide (FA) and
LOHC system based on ethylenediamine-ethanol/N,N′- N,N′-(ethane-1,2-diyl)diformamide (DFA) in 15% and 9%
diacetylethylenediamine (DAE), with a hydrogen storage yields, respectively.
capacity of 5.3 wt % (Scheme 13). However, because methanol Other ruthenium pincer complexes 4 and 6 were also
has one fewer carbon and the dehydrogenative coupling of ED examined for the dehydrogenative coupling of ED and
and methanol can result in the formation of ethylene urea methanol, and neither the intermediate FA nor EU was
(EU), which has a lower molecular weight than DAE, the observed under similar reaction conditions. According to
hydrogen storage capacity of a system based on methanol mechanistic experiments, the dehydrogenative coupling of ED
would be higher (6.45 wt % H2 capacity). Using 1 mol % 9 and and methanol using catalyst 9 proceeds via metal−ligand
4 mol % KOtBu, the hydrogenation of EU under 60 bar cooperation (MLC) and results in the formation of FA and
hydrogen pressure at 170 °C for 7 d in 2 mL dioxane preceded DFA with the release of molecular hydrogen. DFA can be
with complete conversion and the products ED and methanol transformed to FA in the presence of ED and then to EU via
were obtained in 100% selectivity (Scheme 18). Moreover, the dehydrogenation reaction (Scheme 19).
14721 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis


pubs.acs.org/acscatalysis Review

ALCOHOL-BASED LOHC SYSTEMS mol % of 17, the crude reaction mixture was fully
In 2016, Gusev reported an interesting hydrogen storage hydrogenated to EG in 48 h at 40 bar of H2 pressure. Under
system based on ethanol/ethyl acetate couple via the reversible neat conditions and reduced pressure (95 mbar), a large-scale
dehydrogenation/hydrogenation reaction catalyzed by osmi- dehydrogenation of EG (2 mL) at 150 °C for 7 days afforded
um-based pincer complexes.71 The ADC of ethanol using 0.01 94% conversion and 1295 mL (75% H2 yield) of hydrogen.
mol % 14 and 1 mol % NaOEt as a base at 90 °C resulted in Using 0.5 mol % 17 and at 40 bar of H2 pressure, this crude
82% ethanol conversion to ethyl acetate (Scheme 20). mixture was completely converted to EG after 60 h.
The mechanism of this reversible dehydrogenation/hydro-
Scheme 20. Reversible Dehydrogenation/Hydrogenation of genation of EG based on calculations using the density
Ethanol and Ethyl Acetate functional theory has been illustrated in Scheme 22. There are
numerous challenges associated with the acceptorless dehy-
drogenative coupling of EG to 2-hydroxyethyl glycolate
(HEG), including EG chelation with the metal center, which
reduces catalyst activity, hydrogen bonding between the alkoxy
metal complex and neighboring EG, which may obstruct the β-
hydride elimination step, dehydrogenation of HEG to a α-keto
ester, which further decomposes to CO and aldehyde, and the
formation of undesired products with lower HSCs such as
((1,3-dioxolan-2-yl)methanol). Nevertheless, the acridine
ligand backbone in ruthenium complex 17 has unique
characteristics which are responsible for overcoming these
challenges. Moreover, the dehydrogenation of EG to HEG was
endergonic to a small extent, which indicates the feasibility of
reversible dehydrogenation/hydrogenation reactions.
Notably, the reverse hydrogenation reaction was catalyzed by Later, Fujita et al.73 developed a hydrogen carrier system
the dihydride osmium complex 15 at 50 bar of hydrogen based on the formation of γ-butyrolactone by dehydrogenative
pressure. Indeed, after 1 h, relatively higher TONs of 730 were lactonization of 1,4-butanediol and regeneration of 1,4-
observed. butanediol reversely through catalytic hydrogenation reaction.
Milstein and co-workers also described a liquid-to-liquid The dehydrogenative lactonization proceeded successfully
organic hydrogen carrier system based on ethylene glycol under solvent-free conditions with an iridium catalyst 18
(EG), which is cheap, readily available, and renewable.72 With containing a functional 6,6′-dionato-2,2′-bipyridine ligand,
a theoretical hydrogen storage potential of 6.5 wt %, ethylene yielding both hydrogen and γ-butyrolactone in 96% yields
glycol might efficiently store and release hydrogen reversibly after 9 h (Scheme 23). The reverse hydrogenation of γ-
using a ruthenium pincer complex. By utilizing 1 mol % of butyrolactone in the presence of 1 mol % 18 under 8 atm H2
acridine-based PNP ruthenium complex 16 and 2 mol % pressure also proceeded quantitatively, and 1,4-butanediol was
KOtBu, 94% conversion of EG and 54 mL of hydrogen with obtained in 98% yield after 72 h. In addition, triethylamine and
99.65% purity were obtained (Scheme 21). Further, using 1 6,6′-dihydroxy-2,2′-bipyridine were required as additives in the
mol % of dearomatized complex 17, the dehydrogenative hydrogenation reaction.
coupling of EG proceeded efficiently under base-free Another liquid-to-liquid organic hydrogen carrier system
conditions and yielded 97% conversion and 61 mL of based on ruthenium complex catalyzed reversible dehydrogen-
hydrogen (99.59% purity). Moreover, in the presence of 1 ative coupling of ethylene glycol (EG) with ethanol under
solvent and base-free conditions was disclosed by the Milstein
Scheme 21. Reversible Liquid Hydrogen Carrier System group.74 Although numerous LOHC systems based on the
Based on the Acceptorless Dehydrogenative Coupling of amidation reaction of alcohols and amines have been
Ethylene Glycol72 developed, the products obtained are solid at room temper-
ature and require the use of a solvent, which reduces the
hydrogen storage capacity (HSC). Initially, the intermolecular
dehydrogenative esterification of EG with ethanol using 1 mol
% Ru-16 and 1 mol % KOtBu in a solvent mixture of toluene
and DMF at 150 °C for 72 h, resulted in 87% yield of H2 with
greater than 99.85% purity (Scheme 24). Furthermore, under
base-free conditions, the reaction of EG and ethanol with the
stable dearomatized acridine PNP ruthenium complex Ru-17
yielded 80% hydrogen. A 50 bar H2 pressure was used to fully
hydrogenate the ester mixture back to EG and ethanol.
Because dehydrogenation with solvents reduces the system’s
HSC, EG and ethanol were reacted in large scale under
solvent- and base-free conditions resulting in a 71% hydrogen
yield and a 72% EG and ethanol yield from reverse
hydrogenation. In Scheme 25, the proposed catalytic pathway
for the dehydrogenative coupling of EG and ethanol is
provided.
14722 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Scheme 22. A Plausible Mechanism of the Acceptorless Dehydrogenative Coupling of Ethylene Glycol

Scheme 23. Reversible Dehydrogenation (or Lactonization) mostly noble metals-based catalysts are involved especially
of 1,4-Butanediol into γ-Butyrolactone73 ruthenium, iridium with expensive ligand system, replacing
these metals with earth-abundant, and inexpensive 3d
transition metals with cost-effective ligands and heterogeneous
catalytic systems could be a promising approach for
economically viable LOHC in near future. Thus, it requires
more developments and infrastructure for practical applica-
tions.

■ CONCLUSIONS AND FUTURE PERSPECTIVE


With rapid industrialization and population growth, the global
energy demand is continuously increasing. So far, carbon-based
fossil fuels such as oil, coal, and natural gas have met the
majority of energy demands. However, challenges like fossil
fuel depletion and anthropogenic global warming, as well as
their negative effects on human health and the environment,
have prompted a search for alternative energy sources to meet
the modern world’s energy demands. Hydrogen is widely
considered a clean and environment-friendly fuel because of its


high gravimetric energy density and clean-burning properties.
The conventional hydrogen storage technologies, including
LIMITATIONS OF ALCOHOL-BASED LOHC compression and liquefaction, are costly and unsafe.
SYSTEMS Furthermore, the chemical hydrogen storage methods such
The hydrogen economy is the center of attention in several as hydrogen adsorption in metal hydrides, metal clusters,
developed countries. Because of the main concern of climate metal−organic frameworks (MOFs), and nanostructured
change due to CO2 emission and environmental pollutants, materials also suffer from several aspects. Liquid organic
hydrogen will be an ideal replacement for all the conventional hydrogen carriers (LOHCs), which store hydrogen in covalent
energy options. At present, the economic feasibility of an bonds of liquid organic compounds, are quite promising. In
alcohol-based LOHC system with exciting methodology is this Review, we have covered several essential aspects of
challenging to contribute to the hydrogen economy. Because developing novel, cost-effective, and abundantly available
14723 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Scheme 24. Ethylene Glycol and Ethanol-Based Reversible LOHC System74

Scheme 25. Proposed Route for the Dehydrogenative orcid.org/0000-0001-9248-2809; Email: eb.raman@
Coupling of EG and Ethanol iisertirupati.ac.in
Authors
Vinita Yadav − Organic Chemistry Division, CSIR-National
Chemical Laboratory (CSIR-NCL), Pune, Maharashtra
411008, India; Academy of Scientific and Innovative
Research (AcSIR), Ghaziabad 201002, India
Ganesan Sivakumar − Department of Chemistry, Indian
Institute of Science Education and Research (IISER)
Tirupati, Tirupati 517507, India
Virendrakumar Gupta − Polymer Synthesis & Catalysis,
Reliance Research & Development Centre, Reliance Industries
Limited, Navi Mumbai 400701, India; orcid.org/0000-
0002-3896-1451
Complete contact information is available at:
https://pubs.acs.org/10.1021/acscatal.1c03283
LOHC systems for efficient hydrogen storage and transport
applications. LOHCs are very similar to crude oil derivatives at Notes
ambient conditions, thus using existing energy infrastructure. The authors declare no competing financial interest.


Without any doubt, notable developments has been made in
the effectiveness of LOHCs hydrogenation and dehydrogen-
ation process; still, further research and development are ACKNOWLEDGMENTS
needed. E.B. acknowledges funding from Swarnajayanti Fellowship


*
ASSOCIATED CONTENT
sı Supporting Information
(DST/SJF/CSA-04/2019-2020 & SERB/F/5892/2020-2021),
and SERB, India (Grant No: CRG/2018/002480/OC). V.Y.
acknowledges the UGC, India for fellowship. G.S. thanks
The Supporting Information is available free of charge at IISER-Tirupati for fellowship.


https://pubs.acs.org/doi/10.1021/acscatal.1c03283.
Various catalysts used in the catalytic hydrogenation of REFERENCES
CO2 to methanol and price of transition-metals used in
the dehydrogenation reactions (PDF) (1) El-Shafie, M.; Kambara, S.; Hayakawa, Y. Hydrogen Production


Technologies Overview. J. Power Energy Eng. 2019, 7, 107−154.
(2) Singh, S.; Jain, S.; PS, V.; Tiwari, A. K.; Nouni, M. R.; Pandey, J.
AUTHOR INFORMATION K.; Goel, S. Hydrogen: A Sustainable Fuel for Future of the Transport
Corresponding Author Sector. Renewable Sustainable Energy Rev. 2015, 51, 623−633.
Ekambaram Balaraman − Department of Chemistry, Indian (3) Satyapal, S.; Petrovic, J.; Read, C.; Thomas, G.; Ordaz, G. The
Institute of Science Education and Research (IISER) U.S. Department of Energy’s National Hydrogen Storage Project:
Tirupati, Tirupati 517507, India; Academy of Scientific and Progress towards Meeting Hydrogen-Powered Vehicle Requirements.
Innovative Research (AcSIR), Ghaziabad 201002, India; Catal. Today 2007, 120, 246−256.

14724 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

(4) Modisha, P. M.; Ouma, C. N.; Garidzirai, R.; Wasserscheid, P.; (27) Fuel Cells and Hydrogen Joint Undertaking (FCH2 JU)
Bessarabov, D. The Prospect of Hydrogen Storage using Liquid Governing Board. Multi-Annual Work Plan 2014−2020; FCH2 JU:
Organic Hydrogen Carriers. Energy Fuels 2019, 33, 2778−2796. Brussels, Belgium, 2014.
(5) Hosseini, S. E.; Wahid, M. A. Hydrogen Production from (28) U.S. DRIVE Partnership. Target Explanation Document:
Renewable and Sustainable Energy Resources: Promising Green Onboard Hydrogen Storage for Light-Duty Fuel Cell Vehicles; U.S.
Energy Carrier for Clean Development. Renewable Sustainable Energy Department of Energy: Washington, DC, 2014.
Rev. 2016, 57, 850−866. (29) Olah, G. A. Towards Oil Independence Through Renewable
(6) Jones, L. W. Liquid Hydrogen as a Fuel for the Future. Science Methanol Chemistry. Angew. Chem., Int. Ed. 2013, 52, 104−107.
1971, 174, 367−370. (30) Porosoff, M. D.; Yan, B. H.; Chen, J. G. G. Catalytic Reduction
(7) Winsche, W. E.; Hoffman, K. C.; Salzano, F. J. Hydrogen: Its of CO2 by H2 for Synthesis of CO, Methanol and Hydrocarbons:
Future Role in the Nation’s Energy Economy. Science 1973, 180, Challenges and Opportunities. Energy Environ. Sci. 2016, 9, 62−73.
1325−1332. (31) Kar, S.; Kothandaraman, J.; Goeppert, A.; Prakash, G. K. S.
(8) Momirlan, M.; Veziroglu, T. N. The Properties of Hydrogen as Advances in Catalytic Homogeneous Hydrogenation of Carbon
Fuel Tomorrow in Sustainable Energy System for a Cleaner Planet. Dioxide to Methanol. J. CO2 Util. 2018, 23, 212−218.
Int. J. Hydrogen Energy 2005, 30, 795−802. (32) Alberico, E.; Nielsen, M. Towards a Methanol Economy based
(9) Bockris, J. O. M. The Hydrogen Economy: Its History. Int. J. on Homogeneous Catalysis: Methanol to H2 and CO2 to methanol.
Hydrogen Energy 2013, 38, 2579−2588. Chem. Commun. 2015, 51, 6714−6725.
(10) Navarro, R. M.; Sánchez-Sánchez, M. C.; Alvarez-Galvan, M. (33) Teichmann, D.; Arlt, W.; Wasserscheid, P.; Freymann, R. A
C.; del Valle, F.; Fierro, J. L. G. Hydrogen Production from Future Energy Supply based on Liquid Organic Hydrogen Carriers
Renewable Sources: Biomass and Photocatalytic Opportunities. (LOHC). Energy Environ. Sci. 2011, 4, 2767−2773.
Energy Environ. Sci. 2009, 2, 35−54. (34) Teichmann, D.; Arlt, W.; Wasserscheid, P. Liquid Organic
(11) Armaroli, N.; Balzani, V. The Hydrogen Issue. ChemSusChem Hydrogen Carriers as an Efficient Vector for the Transport and
2011, 4, 21−36. Storage of Renewable Energy. Int. J. Hydrogen Energy 2012, 37,
(12) Scott, K. In Electrochemical Methods for Hydrogen Production 18118−18132.
2019, 1, 1−27. (35) Cooper, A. C.; Fowler, D. E.; Scott, A. R.; Abdourazak, A. H.;
(13) Ustolin, F.; Paltrinieri, N.; Berto, F. Loss of Integrity of Cheng, H.; Wilhelm, F. C.; Toseland, B. A.; Campbell, K. M.; Pez, G.
Hydrogen Technologies: A Critical Review. Int. J. Hydrogen Energy P. Hydrogen Storage by Reversible Hydrogenation of Liquid-Phase
2020, 45, 23809−23840. Hydrogen Carriers. Prepr. Pap. Am. Chem. Soc. Div. Fuel Chem. 2005,
(14) Sadaghiani, M. S.; Mehrpooya, M. Introducing and Energy 50, 271−273.
Analysis of a Novel Cryogenic Hydrogen Liquefaction Process (36) Sekine, Y.; Higo, T. Recent Trends on the Dehydrogenation
Configuration. Int. J. Hydrogen Energy 2017, 42, 6033−6050. Catalysis of Liquid Organic Hydrogen Carrier (LOHC): A Review.
(15) Sordakis, K.; Tang, C.; Vogt, L. K.; Junge, H.; Dyson, P. J.; Top. Catal. 2021, 64, 470−480.
Beller, M.; Laurenczy, G. Homogeneous Catalysis for Sustainable (37) Rao, P. C.; Yoon, M. Potential Liquid-Organic Hydrogen
Hydrogen Storage in Formic Acid and Alcohols. Chem. Rev. 2018, Carrier (LOHC) Systems: A Review on Recent Progress. Energies
118, 372−433. 2020, 13, 6040−6062.
(16) Sreedhar, I.; Kamani, K. M.; Kamani, B. M.; Reddy, B. M.; (38) Yamaguchi, R.; Ikeda, C.; Takahashi, Y.; Fujita, K.-I.
Venugopal, A. A Bird’s Eye View on Process and Engineering Aspects Homogeneous Catalytic System for Reversible Dehydrogenation-
of Hydrogen Storage. Renewable Sustainable Energy Rev. 2018, 91, Hydrogenation Reactions of Nitrogen Heterocycles with Reversible
838−860. Interconversion of Catalytic Species. J. Am. Chem. Soc. 2009, 131,
(17) Gianotti, E.; Taillades-Jacquin, M.; Rozière, J.; Jones, D. J. 8410−8412.
High-Purity Hydrogen Generation via Dehydrogenation of Organic (39) Fujita, K.-I.; Tanaka, Y.; Kobayashi, M.; Yamaguchi, R.
Carriers: A Review on the Catalytic Process. ACS Catal. 2018, 8, Homogeneous Perdehydrogenation and Perhydrogenation of Fused
4660−4680. Bicyclic N-Heterocycles Catalyzed by Iridium Complexes Bearing a
(18) He, T.; Pachfule, P.; Wu, H.; Xu, Q.; Chen, P. Hydrogen Functional Bipyridonate Ligand. J. Am. Chem. Soc. 2014, 136, 4829−
Carriers. Nat. Rev. Mater. 2016, 1, 16059−16075. 4832.
(19) Eberle, U.; Felderhoff, M.; Schüth, F. Chemical and Physical (40) Markiewicz, M.; Zhang, Y. Q.; Bösmann, A.; Brückner, N.;
Solutions for Hydrogen Storage. Angew. Chem., Int. Ed. 2009, 48, Thöming, J.; Wasserscheid, P.; Stolte, S. Environmental and Health
6608−6630. Impact Assessment of Liquid Organic Hydrogen Carrier (LOHC)
(20) Dalebrook, A. F.; Gan, W.; Grasemann, M.; Moret, S.; Systems - Challenges and Preliminary Results. Energy Environ. Sci.
Laurenczy, G. Hydrogen Storage: Beyond Conventional Methods. 2015, 8, 1035−1045.
Chem. Commun. 2013, 49, 8735−8751. (41) Faisca Phillips, A. M.; Pombeiro, A. J. L.; Kopylovich, M. N.
(21) Schneemann, A.; White, J. L.; Kang, S.; Jeong, S.; Wan, L. F.; Recent Advances in Cascade Reactions Initiated by Alcohol
Cho, E. S.; Heo, T. W.; Prendergast, D.; Urban, J. J.; Wood, B. C.; Oxidation. ChemCatChem 2017, 9, 217−246.
Allendorf, M. D.; Stavila, V. Nanostructured Metal Hydrides for (42) Barta, K.; Ford, P. C. Catalytic Conversion of Nonfood Woody
Hydrogen Storage. Chem. Rev. 2018, 118, 10775−10839. Biomass Solids to Organic Liquids. Acc. Chem. Res. 2014, 47, 1503−
(22) Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D.-W. Hydrogen 1512.
Storage in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 782− (43) Smith, T. A.; Aplin, R. P.; Maitlis, P. M. The Ruthenium-
835. Catalysed Conversion of Methanol into Methyl Formate. J. Organo-
(23) Garg, N.; Sarkar, A.; Sundararaju, B. Recent Developments on met. Chem. 1985, 291, C13−C14.
Methanol as Liquid Organic Hydrogen Carrier in Transfer Hydro- (44) Yang, L.-C.; Ishida, T.; Yamakawa, T.; Shinoda, S. Mechanistic
genation Reactions. Coord. Chem. Rev. 2021, 433, 213728−213746. Study on Dehydrogenation of Methanol with [RuCl2(PR3)3]-type
(24) Heim, L. E.; Konnerth, H.; Prechtl, M. H. G. Future Catalyst in Homogeneous Solutions. J. Mol. Catal. A: Chem. 1996,
Perspectives for Formaldehyde: Pathways for Reductive Synthesis 108, 87−93.
and Energy Storage. Green Chem. 2017, 19, 2347−2355. (45) Shinoda, S.; Itagaki, H.; Saito, Y. Dehydrogenation of Methanol
(25) Eppinger, J.; Huang, K.-W. Formic Acid as a Hydrogen Energy in the Liquid Phase with a Homogeneous Ruthenium Complex
Carrier. ACS Energy Lett. 2017, 2, 188−195. Catalyst. J. Chem. Soc., Chem. Commun. 1985, 860−861.
(26) Olah, G. A. Beyond Oil and Gas: The Methanol Economy. (46) Dobson, A.; Robinson, S. D. Complexes of the Platinum
Angew. Chem., Int. Ed. 2005, 44, 2636−2639. Metals. 7. Homogeneous Ruthenium and Osmium Catalysts for the

14725 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726
ACS Catalysis pubs.acs.org/acscatalysis Review

Dehydrogenation of Primary and Secondary Alcohols. Inorg. Chem. Peptide Formation and Hydrogenation. Nat. Commun. 2015, 6,
1977, 16, 137−142. 6859−6865.
(47) Itagaki, H.; Shinoda, S.; Saito, Y. Liquid-Phase Dehydrogen- (66) Hu, P.; Ben-David, Y.; Milstein, D. Rechargeable Hydrogen
ation of Methanol with Homogeneous Ruthenium Complex Catalysts. Storage System Based on the Dehydrogenative Coupling of
Bull. Chem. Soc. Jpn. 1988, 61, 2291−2294. Ethylenediamine with Ethanol. Angew. Chem. 2016, 128, 1073−1076.
(48) Morton, D.; Cole-Hamilton, D. J. Molecular Hydrogen (67) Kumar, A.; Janes, T.; Espinosa-Jalapa, N. A.; Milstein, D.
Complexes in Catalysis: Highly Efficient Hydrogen Production from Selective Hydrogenation of Cyclic Imides to Diols and Amines and Its
Alcoholic Substrates Catalysed by Ruthenium Complexes. J. Chem. Application in the Development of a Liquid Organic Hydrogen
Soc., Chem. Commun. 1988, 1154−1156. Carrier. J. Am. Chem. Soc. 2018, 140, 7453−7457.
(49) Morton, D.; Cole-Hamilton, D. J.; Utuk, I.; Paneque-Sosa, M.; (68) Kothandaraman, J.; Kar, S.; Sen, R.; Goeppert, A.; Olah, G. A.;
Lopez-Poveda, M. Hydrogen Production from Ethanol Catalysed by Prakash, G. K. S. Efficient Reversible Hydrogen Carrier System Based
Group 8 Metal Complexes. J. Chem. Soc., Dalton Trans. 1989, 489− on Amine Reforming of Methanol. J. Am. Chem. Soc. 2017, 139,
495. 2549−2552.
(50) Shinoda, S.; Yamakawa, T. One-step Formation of Methyl (69) Shao, Z.; Li, Y.; Liu, C.; Ai, W.; Luo, S.-P.; Liu, Q. Reversible
Acetate with Methanol used as the Sole Source and Catalysis by RuII- Interconversion between Methanol-Diamine and Diamide for Hydro-
SnII Cluster Complexes. J. Chem. Soc., Chem. Commun. 1990, 1511− gen Storage based on Manganese Catalyzed (De)Hydrogenation. Nat.
1512. Commun. 2020, 11, 591−597.
(51) Shinoda, S.; Ohnishi, T.; Yamakawa, T. Single-step Synthesis of (70) Xie, Y.; Hu, P.; Ben-David, Y.; Milstein, D. A Reversible Liquid
Acetic Acid (Methyl Acetate) from Methanol alone by Ru(II)-Sn(II) Organic Hydrogen Carrier System Based on Methanol-Ethylenedi-
amine and Ethylene Urea. Angew. Chem., Int. Ed. 2019, 58, 5105−
Hetero-Bimetallic Catalysts. Catal. Surv. Jpn. 1997, 1, 25−34.
5109.
(52) Robles-Dutenhefner, P. A.; Moura, E. M.; Gama, G. J.; Siebald,
(71) Gusev, D. G. Dehydrogenative Coupling of Ethanol and Ester
H. G. L.; Gusevskaya, E. V. Synthesis of Methyl Acetate from
Hydrogenation Catalyzed by Pincer-Type YNP Complexes. ACS
Methanol Catalyzed by [(η5-C5H5)(phosphine)2RuX] and [(η5-
Catal. 2016, 6, 6967−6981.
C5H5)(phosphine)2Ru(SnX3)] (X = F, Cl, Br): Ligand Effect. J. (72) Zou, Y.-Q.; von Wolff, N.; Anaby, A.; Xie, Y.; Milstein, D.
Mol. Catal. A: Chem. 2000, 164, 39−47. Ethylene Glycol as an Efficient and Reversible Liquid Organic
(53) Shimbayashi, T.; Fujita, K.-I. Metal-Catalyzed Hydrogenation Hydrogen Carrier. Nature Catalysis 2019, 2, 415−422.
and Dehydrogenation Reactions for Efficient Hydrogen Storage. (73) Onoda, M.; Nagano, Y.; Fujita, K. Iridium-Catalyzed
Tetrahedron 2020, 76, 130946−130973. Dehydrogenative Lactonization of 1,4-Butanediol and Reversal
(54) Sponholz, P.; Mellmann, D.; Cordes, C.; Alsabeh, P. G.; Li, B.; Hydrogenation: New Hydrogen Storage System using Cheap Organic
Li, Y.; Nielsen, M.; Junge, H.; Dixneuf, P.; Beller, M. Efficient and Resources. Int. J. Hydrogen Energy 2019, 44, 28514−28520.
Selective Hydrogen Generation from Bioethanol using Ruthenium (74) Zhou, Q.-Q.; Zou, Y.-Q.; Ben-David, Y.; Milstein, D. A
Pincer-type Complexes. ChemSusChem 2014, 7, 2419−2422. Reversible Liquid-to-Liquid Organic Hydrogen Carrier System Based
(55) Rybak, W. K.; Ziółkowski, J. J. Dehydrogenation of Alcohols on Ethylene Glycol and Ethanol. Chem. - Eur. J. 2020, 26, 15487−
Catalyzed by Polystyrene-Supported Ruthenium Complexes. J. Mol. 15490.
Catal. 1981, 11, 365−370.
(56) Jung, C. W.; Garrou, P. E. Dehydrogenation of Alcohols and
Hydrogenation of Aldehydes using Homogeneous Ruthenium
Catalysts. Organometallics 1982, 1, 658−666.
(57) Ligthart, G. B. W. L.; Meijer, R. H.; Donners, M. P. J.;
Meuldijk, J.; Vekemans, J. A. J. M.; Hulshof, L. A. Highly Sustainable
Catalytic Dehydrogenation of Alcohols with Evolution of Hydrogen
Gas. Tetrahedron Lett. 2003, 44, 1507−1509.
(58) Junge, H.; Loges, B.; Beller, M. Novel Improved Ruthenium
Catalysts for the Generation of Hydrogen from Alcohols. Chem.
Commun. 2007, 5, 522−524.
(59) 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. 2011, 50, 9593−9597.
(60) Spasyuk, D.; Gusev, D. G. Acceptorless Dehydrogenative
Coupling of Ethanol and Hydrogenation of Esters and Imines.
Organometallics 2012, 31, 5239−5242.
(61) Nielsen, M.; Junge, H.; Kammer, A.; Beller, M. Towards a
Green Process for Bulk-Scale Synthesis of Ethyl Acetate: Efficient
Acceptorless Dehydrogenation of Ethanol. Angew. Chem., Int. Ed.
2012, 51, 5711−5713.
(62) Hu, P.; Ben-David, Y.; Milstein, D. General Synthesis of Amino
Acid Salts from Amino Alcohols and Basic Water Liberating H2. J. Am.
Chem. Soc. 2016, 138, 6143−6146.
(63) Wang, Y.; Wang, M.; Li, Y.; Liu, Q. Homogeneous Manganese-
Catalyzed Hydrogenation and Dehydrogenation Reactions. Chem.
2021, 7, 1180−1223.
(64) Chakraborty, S.; Lagaditis, P. O.; Förster, M.; Bielinski, E. A.;
Hazari, N.; Holthausen, M. C.; Jones, W. D.; Schneider, S. Well-
Defined Iron Catalysts for the Acceptorless Reversible Dehydrogen-
ation-Hydrogenation of Alcohols and Ketones. ACS Catal. 2014, 4,
3994−4003.
(65) Hu, P.; Fogler, E.; Diskin-Posner, Y.; Iron, M. A.; Milstein, D. A
Novel Liquid Organic Hydrogen Carrier System based on Catalytic

14726 https://doi.org/10.1021/acscatal.1c03283
ACS Catal. 2021, 11, 14712−14726

You might also like