ChemSusChem - 2021 - Guo - Formic Acid as a Potential on‐Board Hydrogen Storage Method Development of Homogeneous Noble

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Reviews

ChemSusChem doi.org/10.1002/cssc.202100602

Formic Acid as a Potential On-Board Hydrogen Storage


Method: Development of Homogeneous Noble Metal
Catalysts for Dehydrogenation Reactions
Jian Guo,[a] Chengkai K. Yin,[b] Dulin L. Zhong,[a] Yilin L. Wang,[a] Tiangui Qi,[a] Guihua H. Liu,[a]
Leiting T. Shen,[a] Qiusheng S. Zhou,[a] Zhihong H. Peng,[a] Hong Yao,*[b] and Xiaobin B. Li*[a]

ChemSusChem 2021, 14, 2655 – 2681 2655 © 2021 Wiley-VCH GmbH


1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Hydrogen can be used as an energy carrier for renewable carbon dioxide through catalysis, has significant potential for
energy to overcome the deficiency of its intrinsically intermit- practical application. Historic developments and recent exam-
tent supply. One of the most promising application of hydrogen ples of homogeneous noble metal catalysts for FA dehydrogen-
energy is on-board hydrogen fuel cells. However, the lack of a ation are covered, and the catalysts are classified based on their
safe, efficient, convenient, and low-cost storage and trans- ligand types. The Review primarily focuses on the
portation method for hydrogen limits their application. The structure function relationship between the ligands and their
feasibility of mainstream hydrogen storage techniques for reactivity and aims to provide suggestions for designing new
application in vehicles is briefly discussed in this Review. Formic and efficient catalysts for H2 generation from FA.
acid (FA), which can reversibly be converted into hydrogen and

Introduction Hydrogen energy generated from renewable resources can


be a perpetual solution to global energy and environmental
The rapid development of human society has triggered critical problems.[10–14] Hydrogen has great advantages as an energy
energy and environmental problems. The worldwide total carrier. The storage period of hydrogen is longer than that of
energy consumption increased 2.3 times, from 4242 Mtoes in electric energy, and it can be used as a long-term energy
1971 to 9717 Mtoes in 2017 (1 Mtoe � 42 GJ). The energy storage method to store renewable electric energy and increase
consumption of most sectors was relatively stable, although the energy flexibility.[15] Hydrogen has great environmental compat-
energy consumption contribution of the transportation sector ibility, as water is formed as the only by-product. The mass
increased significantly from 23 % in 1971 to 29 % in 2017. energy density of hydrogen is 142 KJ/g, approximately three
During this period, the world’s total primary energy supply times that of petrol.[16,17] Hydrogen can be utilized in almost all
increased by a factor of 2.5, from 5519 Mtoes to 13972 Mtoes sectors, including transportation, home energy supply, and
(Figure 1).[1] Numerically, the growth rate of energy supply commercial fields.[18]
should match that of energy consumption. Fossil fuels, On-board hydrogen fuel cells are an attractive utilization of
including coal, oil, and natural gas, account for more than 80 % hydrogen. Such fuel cells are highly efficient and present an
of the present-day energy supply system.[2] With increasing eco-friendly technology for energy use.[19–23] In transportation,
energy consumption and depleting fossil fuel sources, the the efficiency of hydrogen fuel cell engines is as high as 65 %,
present energy supply system will not satisfy the persistently compared to 25 % for gasoline powered engines.[24] When the
growing energy demand. Meanwhile, the heavy use of fossil heat generated by a hydrogen fuel cell is utilized in combined
fuels has caused rapid growth in carbon dioxide emissions. The heat and power systems, the overall efficiency can exceed
average annual increase of carbon dioxide in the atmosphere 85 %.[24] Hydrogen adoption can start with passenger cars and
was 1.28 ppm from 1970 to 1979 and increased to 2.40 ppm buses.[15] General Motors in the United States has evaluated 75
from 2010 to 2019.[3] High levels of carbon dioxide in the fuel pathways for vehicles. Hydrogen fuel-cell vehicles exhibit
atmosphere causes the greenhouse effect.[4] Increasing carbon low overall energy consumption and low carbon dioxide
dioxide emissions lead to high global temperatures causing emissions.[25]
several ecological problems.[5] Based on the aforementioned Although hydrogen has a high mass energy density, its
problems, researchers have started focusing on eco-friendly volumetric energy density is not competitive for vehicles due to
renewable resources. its gaseous state at ambient temperatures and pressures.[17] On-
Renewable resources, including solar, wind, tidal, hydro- board hydrogen storage systems, which should not occupy
power, biomass, and geothermal resources,[6] are clean, waste- considerable space, require acceptable mass and volumetric
free, recyclable and have enormous developing potential. energy densities. The materials and equipment should have
However, the intrinsic intermittency of renewable energy sufficient durability and safety. Storage should be reversible,
hinders its large-scale use.[7,8] The missing parts of a sustainable and the release of hydrogen should display fast kinetics under
energy system are the energy storage program (used to store mild conditions. The overall cost should be competitive; ideally,
renewable energy when it is not being produced) and energy the storage and transportation equipment should be compat-
carrier (an alternate to gasoline or other fossil-derived energy ible with existing installations. These goals place high technical
carriers). Hydrogen is considered the most generic energy requirements on hydrogen storage systems. Despite significant
storage system and the most promising energy carrier.[9] research, none of the existed H2 storage methods meet all the
aforementioned requirements.[26]
[a] Dr. J. Guo, D. L. Zhong, Dr. Y. L. Wang, T. Qi, Prof. Dr. G. H. Liu,
Prof. Dr. L. T. Shen, Prof. Dr. Q. S. Zhou, Prof. Dr. Z. H. Peng, Prof. Dr. X. B. Li
School of Metallurgy and Environment, Central South University Hydrogen Storage Techniques
932 Lushan Road, Changsha city, Hunan Province, 410083 (P. R. China)
E-mail: x.b.li@csu.edu.cn
[b] Dr. C. K. Yin, H. Yao High pressures increase the volumetric energy density. How-
Hangzhou Katal Catalyst & Metal Material Stock Co., Ltd. ever, high-pressure H2 storage tanks are invariably very heavy,
7 Kang Qiao Road, Gong Shu District, Hang Zhou, Zhejiang Province,
310015 (P. R. China) which lowers the overall mass energy density.[27] Moreover, they
E-mail: kd@katal.com.cn present considerable risks, as compression is a dangerous and

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2656 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Figure 1. a,b) World total energy consumption by sector in 1971 (a) and 2017 (b). c,d) World total energy supply by fuels in 1971 (c) and 2017 (d). Note: Peat
and oil shale are aggregated with coal. Source: IEA(2019) World energy balance Overview. All rights reserved.

complicated process.[28] Liquified hydrogen storage is a viable storage containers require multi-layer thermal insulation materi-
option that is widely applied in aerospace industry. Liquid als and have a cylindrical design,[29] which is not compatible
hydrogen is an ideal aviation fuel owing to its high mass and with the construction of common passenger cars. Evaporation
volumetric energy densities. However, low-temperature H2 losses during hydrogenation and storage, as well as the high

Hong Yao, senior engineer, obtained his B.Sc. Xaobin Li received his B.Eng. and M.Eng. in
at Central South University, P.R. China, in nonferrous metallurgy from Central South
1982. From 1996 to 2004, he was director of Institute of Mining and Metallurgy, P.R. China,
the Precious Metals Research Office of Zhe- in 1983 and 1985 respectively, and his Ph.D. in
jiang Institute of Metallurgy and in-service Metallurgical Physicochemistry from North-
graduate student at Zhejiang University of eastern University, P.R. China, in 2000. In 1997,
Technology. From 2005 to 2014, he was the he was awarded the Royal Society Fellowship
Executive Director and general manager of to do research at Birmingham University(U.K.)
Hangzhou Kaida Catalyst & Metal Material as a senior research fellow for one year.
Stock Co., Ltd. Since 2014 he has been chair- Presently, Prof. Li is professor and academic
man and general manager of Hangzhou Kaida head of the Alkaline Process Metallurgy
Catalyst & Metal Material Stock Co., Ltd. His Research Center at Central South University.
research interests are the design and synthesis His research interests include the theory and
of noble metal catalysts for exhaust gas treat- technique of alkaline process metallurgy,
ment, hydrogen fuel cells, and H2 production comprehensive utilization of resources and
from small organic molecules. environmental protection, and preparation of
alumina catalyst supports and noble metal
catalysts.

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2657 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

energy consumption of the H2 liquefaction process, have led such as cyclohexane and benzene systems with 7.2 wt% or
almost all of the world’s large car manufacturers to believe that 56 g/L H2, was presented as a hydrogen storage method.
liquid hydrogen is unrealistic for on-board utilization.[30] However, the release of hydrogen at temperatures above 200 °C
H2 can be stored in porous materials through van der Waals is required for most alkane LOHCs.[45] N-heteroaromatic com-
interactions, electrostatic interactions and orbital interactions pounds are also candidates for H2 storage. The presence of N
with their inner surfaces. Rapid kinetics and reversibility are the atoms improves dehydrogenation kinetics and thermodynam-
primary advantages of H2 storage by physisorption.[31] Tradi- ics. However, the thermal stability is decreased, leading to
tional zeolites, activated carbon, and new carbon-containing undesired reactions.[43] Methanol is easy to handle and contains
nanomaterials, such as metal–organic frameworks (MOFs), have 12.6 wt% hydrogen. The aqueous -phase reforming of methanol
been widely investigated. Theoretical calculations and experi- is a state-of-the-art technology for methanol dehydrogenation.
ments show that zeolite capacity, intrinsically limited by geo- However, the reaction occurs at high temperature (over 200 °C)
metric constraints, is at best 2.86 wt%, which is lower than that and high pressures (20–25 MPa) and suffers from low efficiency
required by mobile application.[32] Activated carbon displays a (40 %) and undesired CO production.[45] Nielsen and co-
high capacity at low temperature, but is not competitive at workers[46] developed Ru complexes for low-temperature aque-
ambient conditions.[33] In the late 1990s, many carbon nano- ous-phase methanol dehydrogenation that significantly lowered
materials with a high H2 capacity were reported, but the the reaction temperature. However, the reaction still requires
capacities were shown to be incorrect afterwards.[34,35] With an further development. Recently, FA has been regarded as one of
understanding of adsorption mechanics and synthesis of new the most promising LOHCs. Though FA has only 4.4 wt% H2, the
carbon nanomaterials, carbon nanomaterials have the potential high density of FA (1.22 g cm 3) results in a volumetric H2
to be hydrogen storage materials in the future.[36] MOFs are a content of 53 g H2/L, which is equivalent to 1.77 kWh L 1. This
combination of organic linkers and inorganic nodes. They exceeds the current industry standard (i. e., the Toyota Mirai
enable the design and synthesis of porous materials with 70 MPa H2 storage tank with 1.4 kWh L 1).[47] As a commonly
optimized properties, such as pore volume, surface area, and H2 used industrial reagent, FA is nonflammable, non-toxic and easy
adsorption coordination sites.[37] The flexible features of some to handle at ambient conditions.[48] FA dehydrogenation to
MOFs enable an increase in their capacity.[38] An MOF with a H2 carbon dioxide and its reverse process, carbon dioxide hydro-
adsorption capacity of 9.95 wt% H2 at 77 K[39] have been genation, can be achieved using a suitable catalyst at mild
reported. However, the capacity decreased significantly at temperatures, resulting in a “carbon neutral cycle”.[49] The two
ambient temperatures. The strong temperature dependency of possible pathways of FA decomposition are given by Equa-
H2 physisorption on porous materials and, consequently, on tions (1) and (2):[50]
their capacities is the main obstacle for the application of
porous materials for H2 storage.[40] Furthermore, porous materi- HCOOH ! CO2 þ H2
als tend to have low material densities and, therefore, medium 1 (1)
DG ¼ -48:4 kJ mol
volumetric H2 capacities.[34]
Solid-state H2 storage systems include alloys and metal HCOOH ! CO þ H2 O
1 (2)
hydrides. Typical H2 storage alloys have the form AB5 (e. g., DG ¼ -48:4 kJ mol
LaNi5) or AB2, where A = Ti, Zr, or Mg and B = V, Cr, Fe, or Mn.
However, their hydrogen equilibrium pressures at room temper- Equation (1) represents dehydrogenation to H2 and CO2,
ature are inadequate for fuel cell applications.[30] Metal hydrides whereas Equation (2) represents dehydration to H2O and CO.
can be mainly divided into alanates, borohydrides and amides. The reactions depend on the catalysts used, temperature, and
They have attractive volumetric and gravimetric H2 capacities, FA concentration.[51] CO leads to the deactivation of fuel cell
but suffer from slow kinetics, insufficient reversible capacity, the electrodes. Moreover, the temperature resulting from fuel cell
undesirable evolution of byproducts, and thermal management ‘waste heat’ generation is usually in the range of 80–90 °C.[52]
during recharging.[41] Despite the aforementioned problems, the Therefore, the development of low-temperature, highly efficient
potential development of hydride materials that meet the and FA-selective dehydrogenation catalysts is crucial for the
specifications of the Department of Energy is high.[42] application of FA as a hydrogen energy carrier. Another concern
Liquid organic hydrogen carriers (LOHCs) have attracted of FA dehydrogenation is the generation of CO2 as a by-
increasing attention in recent years. LOHC systems comprise a product.[53] Ideally, CO2, which acts as a H2 carrier, should be
pair of hydrogen-lean and hydrogen-rich organic compounds captured and reused. The use of water as a solvent can be
that store hydrogen through repeated catalytic hydrogenation helpful in recycling CO2 in aqueous medium,[26,54] although the
and dehydrogenation cycles.[43] LOHCs have a significantly recycling process is yet to be achieved.[55] Some researchers
higher volumetric energy density than H2. In addition, LOHCs turn their attention to formates. Papp and Joó[56] presented a
exhibit properties similar to those of crude oil-based fuels, rechargeable H2 storage device that employed the hydro-
enabling the stepwise adaption of existing liquid-fuel genation of bicarbonate and decomposition of formate in
infrastructures.[44] LOHCs are expected to make a breakthrough aqueous solution using Ru-based catalysts. Moreover, the
in the field of H2 storage. device could start a new cycle without isolation of bicarbonate
The reversible catalytic dehydrogenation of cycloalkanes or formate. Beller and co-workers[55] reported that the CO2
and hydrogenation of the corresponding aromatic compounds, content of gas produced from NaHCO2 when using in situ Ru

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2658 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

catalysts was much lower than that from other formates. Later, research progress of FA as a liquid organic hydrogen carrier.
iridium complex containing N-heterocyclic carbene (NHC) The areas of research include the reversible conversion of
ligand[57] and Ru pincer complex[58] were found to be effective carbon dioxide and FA,[104–107] research progress on catalysts
for formate dehydrogenation. Cesium formates were also with different metal centers and various solvent
investigated owing to their higher H2 density compared to their systems,[70,71,108–111] ligand types and a detailed description of the
sodium and potassium analogs. This was attributed to a higher ligand effect on the catalytic reactivity,[112–114] reaction
solubility in water.[59] Separation techniques can also be used to mechanism,[115] and technical issues.[69,116,117] However, we be-
treat CO2. Kawanami and co-workers[60,61] succeeded in separat- lieve that a comprehensive overview of the effects of ligand
ing high-pressure H2 from the gas produced from FA (H2/CO2 = perturbations on catalytic activity could offer new sights into
1 : 1). They could get well purified H2 gas (96 mol%) and CO2 the design of new catalysts. Inspired by Wang’ work,[118] we
liquid (> 99 mol%) by changing the physical state of the gas categorized noble metal homogeneous catalysts into catalysts
mixture at low-temperatures while maintaining a high- with phosphine ligands, catalysts with half-sandwich ligands,
pressure.[53] However, the separation temperature should be catalysts with chelating ligands and other catalysts, owing to
further improved for practical applications. the similarity of the structure-effect relationship between ligand
To date, multitudinous heterogeneous[51,62–65] and and activity. A detailed description of their evolution is
homogeneous[66,67] catalysts have been used for FA dehydrogen- provided. The catalytic system of ionic liquids and non-noble
ation. Heterogeneous catalysts are easy to separate and recycle. metal catalysts are not covered here. We aim to elucidate
However, homogeneous catalysts display better performance in effects of ligand modifications on activities for researchers and
terms of activity and selectivity. promote the development of new catalysts.

Catalysts with phosphine ligands


Homogeneous Catalysts for FA
Dehydrogenation Coffey[68] studied a series of transitional metal catalysts with
phosphine ligands. Depending on the metal center, platinum
The homogeneous catalytic dehydrogenation of FA was first complexes with the general formula [PtCl(PR3)2] displayed
reported in 1967. Coffey[68] reported that FA could be decom- better activity than their nickel and palladium analogs. Ru
posed into hydrogen and carbon dioxide (approximately 1 : 1) complexes with chelating bisphosphine ligands were superior
using phosphine-stabilized complexes of transition metals in to the corresponding osmium complexes. Iridium complexes
acetic acid under reflux. However, it was not until 2008 that the with monodentate phosphine ligands exhibited high activities,
potential application of FA as a liquid organic hydrogen carrier with [IrH3(PPh3)3] displaying the highest activity. The catalysts
was highlighted by Laurenczy and co-workers[66] and Beller and formed complexes containing phosphine, carbonyl, hydrido,
co-workers[67] independently, resuming intensive and acetate ligands during the catalytic reaction. In 1975,
investigations.[26] With continuous development in recent years, several carbonyl-free phosphine complexes of iridium(I), and
homogeneous catalysts have achieved high activity and iridium(III), and rhodium(I) were tested for FA dehydrogenation
selectivity under mild conditions. by Yurtchenko and Anikeenko.[119] The general formula for these
Homogeneous catalysts with noble metal centers, such as Ir, catalysts is [MLxCOy(PR3)z] (M = Ir, Rh; L = Cl, Br, I; x, y, and z vary
Ru and Rh, were found to be effective for H2 generation from with the metal valence state and catalyst structure). For the
FA.[26,69–71] Non-noble metal-based catalyst are attracting consid- same types of compounds, the reaction rate increased in the
erable interest because of their low cost and high availability.[72] order Cl < Br < I. The reaction rate decreased with increasing π-
In 2010, Beller, Ludwig, and co-workers[73] reported the first acceptor ability of the phosphine (PPh3 < PBu3) and depended
iron-based complexes for the catalysis of FA dehydrogenation. on the inherent properties and coordination sites of the ligands.
In 2011, Milstein at al.[74] achieved CO2 hydrogenation to FA For Ir3Cl3L3, the cis- and trans-isomers differed significantly in
using iron complexes containing PNP pincer ligands. Subse- activity. However, as the homogeneous catalytic dehydrogen-
quent intensive studies in the aera have been conducted. ation of FA was not yet a popular topic, many reactions were
Catalytic systems containing iron,[73,75–83] cobalt,[84] discovered during the progress of other scientific studies. In
[85,86]
molybdenum, manganese, magnesium, aluminum,[88,89]
[72] [87]
1979, during an investigation of the reactions of organic
and boron have been investigated. Catalysts with rhenium[91]
[90]
compounds with hydroxy and RhI complexes, Strauss and co-
and osmium[92] metal centers, as well as cell enzymes,[93] were workers[120] found that FA cleaved the rhodium-σ-carbon bond
also explored in this field. However, the state-of-the-art catalysts of [Rh(C6H4PPh2)(PPh3)2] and subsequently formed carbon
for FA dehydrogenation are still those with noble metal dioxide and hydrogen. The decomposition rate was comparable
centers,[72] including Ir,[94–97] Ru,[66,67,98] and Rh.[99] Ionic liquids to that, reported at that time, achieved using group VIII metal
have also been employed, and results have shown that they catalysts. In 1986, The Khai and Arcelli[121] found that
could be helpful in the conversion of FA to H2/CO2 under mild [RuCl2(PPh3)3] could induce FA to decompose into H2 and CO2 in
conditions.[100–103] the presence of triethylamine at ambient temperatures. How-
Researchers have published reviews[26,49] on homogeneous ever, because their research topic was the selective reduction of
catalytic hydrogen production from FA and have described the nitroaromatic compounds, they regarded [RuCl2(PPh3)3]/

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2659 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

HCOOH/Et3N systems as hydrogen sources for hydrogenation.


Until 1988, Puddephat and co-workers[98] made significant
progress in the field. They reported that a binuclear ruthenium
complex [Ru2(μ-CO)(CO)4(μ-dppm)2] (dppm = Ph2PCH2PPh2) cat-
alyzed FA dehydrogenation in acetone at 20 °C, achieving a
turnover frequency (TOF) of 70 h 1, which is considerably higher
than that reported for mononuclear [RuHBrCO(PEt2Ph)3] with a
TOF of 4 h 1 in acetic acid under reflux at 117 °C. Further
mechanism researches were published in 2000, with several
intermediates identified.[122]
In 2008, Beller and co-workers[67] investigated the effects of
different precursors and amines on FA dehydrogenation in FA/
amine adducts. [RhCl3·xH2O] and [RuCl3·xH2O] were almost
inactive, whereas the reactivity of [{RuCl2(p-cymene)}2] increased
with increasing carbon chain length for a series of homologous
aliphatic dimethyl amines. The highest TOF of 2688 h 1 was
obtained with [RuCl2(PPh3)3] in FA/NEt3 at a low temperature of
40 °C. Moreover, the produced H2 could be directly used by
fuel-cells, achieving a similar power density as that of a 1 : 1
mixture of commercial H2 and CO2. In the same year, Laurenczy
and co-workers[66] reported aqueous FA dehydrogenation using
[Ru(H2O)6](tos) (tos = toluene-4-sulfonate) and meta-trisulfo-
Figure 2. Water-soluble sulfonated phosphine ligands L1–L9 reported by
nated triphenylphosphine at 100 °C in a solution of HCOOH/ Laurenczy and co-workers.[124]
HCOONa (9 : 1, 4 M). Moreover, the generated hydrogen was
suitable for all types of fuel cells. Based on their respective
research results, the groups of Beller and Laurenczy independ-
ently proposed the liquid organic hydrogen carrier using Later, Laurenczy and co-workers[125] combined ruthenium ion
FA.[66,67] This initiated a new wave of research. with a group of oligocationic ammoniomethyl-substituted
Laurenczy and co-workers performed various studies on Ru triarylphosphines to achieve FA dehydrogenation (Figure 3).
precursors and phosphines. They found that the activity and The steric demand of these ligands increased in the order L10 <
durability of the catalysts were affected by the ratio of ligand to L11 < L12 and L13 < L14 < L15. To evaluate the electronic
precursor.[123] One equivalent of trisulfonated triphenylphos- properties of the ligands, phosphine selenide analogues were
phine (TPPTS) was not sufficient to stabilize catalytically active synthesized and analyzed by 31P NMR spectroscopy. The σ-
species. However, a system with three equivalent TPPTs was donating strength and basicity of the ligands increased as the
1
stable, but displayed a reduced reaction rate compared to a JP,Se coupling constant decreased.
system with two TPPTs equivalents, indicating an active ligand Experiments showed that FA decomposed faster when
species equivalent to two TPPTs. No activity loss was observed using Ru precursors and ligands with a smaller 1JP,Se, indicating
for the system with [Ru(H2O)6]2 + and two TPPTs equivalent in a that strong σ-donating phosphines promoted the dehydrogen-
continuous hydrogen production experiment after 90 h of use ation reactions. Under the same conditions, systems with
over a period of one month (turnover number (TON) exceeded positively charged ligands (L10–L13, Figure 3) displayed higher
4000 cycles). A reaction mechanism with two competitive activities. This was attributed to the faster coordination of
pathways, in which [RuH(tppts)2(H2O)3] + was an important negatively charged species. The high TOF of 1430 h 1 obtained
intermediate, was proposed based on multinuclear nuclear using L13 was attributed to a balance of adequate σ-donating
magnetic resonance (NMR) spectroscopy. RuCl3 with different strength, desirable steric properties, high hydrophilicity, and
water-soluble sulfonated phosphine ligands were screened in positive valance. In 2014, Laurenczy, Hapiot, Gonsalvi and co-
FA dehydrogenation (Figure 2).[124] The basicity of the ligands workers reported a study on RuCl3·xH2O and different aromatic
was demonstrated by 31P NMR chemical shifts. L6 was the most phosphines containing sulfonate groups (Figure 4).[126] The
basic phosphine and L7 the least basic. L6 and L7 were system decomposed FA into H2 and CO2. At a Ru/L ratio of 2 : 1,
unstable, whereas L1–L5 displayed no activity loss during 20 L17 gave the highest final TOF among the monophosphines of
cycles (Figure 2). 1668 h 1. L16 and L17 were more efficient than their analogues
The position and number of sulfonato groups substantially mTPPMs and mTPPDs (mTPPDs = meta-disulfonated triphenyl-
influenced the solubility and steric effects. L9 displayed better phosphine), which was consistent with their basicities. However,
solubility and activity than L8 (Figure 2), and the solubility of the more basic analogs L21 and L22, containing cycloalkyl
mTPPTs-Na3 was much higher than mTPPMs Na (mTPPMs = groups, displayed lower activity, poor water solubility, and
meta-monosulfonated triphenylphosphine). The high decompo- different steric effects (Figure 4), which are likely the main
sition rate achieved using L1 and L2 was attributed to a factors affecting the performance of these systems. The results
combination of their basicities, steric effects, and solubilities. obtained for diphosphines demonstrated that the presence of

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2660 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Figure 3. Oligocationic triarylphosphine ligands with ammoniomethyl substituents Reaction conditions: Catalysts formed in situ with RuCl3·xH2O (28 mm) and
phosphine (2 equiv), 9 : 1 HCOOH/HCOONa (10 m), 90 °C. TOF was calculated for the fifth cycle of each catalyst.[125]

longer carbon chains reduced the reaction rate. The systems occurred rapidly, whereas [{RuCl2(benzene)}2] with 20 equiv-
with diphosphines achieved higher TONs and lower TOFs than alents of PPh3 maintained its activity for 6 h. In situ catalysts
the systems with monophosphines. The balance between generated from RuCl3·xH2O and three equivalent PPh3 were
activity and durability should also be considered when applying 23 % less active than those for [RuCl2(PPh3)3], indicating the
this type of system. importance of forming catalytically active species. PPh3 and 1,2-
Aside from developing new ligands, Laurenczy and co- bis(diphenylphosphino)butane (dppe) exhibited the best per-
workers performed many studies to promote the practical formances among the tested monodentate and bidentate
application of these catalytic system. They determined the ligands. The optimized catalyst system of N,N’-dimethyl-n-hexyl-
relaxation time of species (i. e., HCO3 /CO32 /CO2) in the solution amine with an in situ generated catalyst based on
in CO2-FA cycles from in situ 1H and 13C NMR spectrum.[127] [RuCl2(benzene)]2 and 6 equivalents of dppe achieved a TON of
These data were used to determine variations with pH and 260000 with an average TOF of nearly 900 h 1.[131] Beller and co-
pressure. More importantly, the data can aid in assessment of workers[132] studied the influence of additives on the Ru-
new catalysts for FA dehydrogenation. To facilitate recycling of catalyzed evolution of hydrogen from FA and found that the
the catalysts, Laurenczy and co-workers studied the immobiliza- reaction rate could be accelerated in the presence of amidines
tion of a homogeneous catalytic system of RuCl3 and TPPTs.[128] and halide additives. Later, they presented the first catalytic FA
They also applied Pt/Al2O3 to remove CO traces from the H2/CO2 dehydrogenation reaction promoted by light. The catalytic
mixture to meet the proton-exchange membrane fuel cell- activity of the Ru precursor and aryl phosphines was signifi-
tolerant levels.[129] cantly enhanced by irradiation, achieving a reaction rate almost
In 2008, Beller and co-workers[130] first reported the effect of twice that in the dark. Moreover, in a continuous experiment,
different phosphine ligands on catalytic H2 production from FA/ the production of H2 could be controlled by irradiation. Further
amine adducts using Ru precursors. In the absence of ligands, NMR studies proved that light assisted the activation of the
RuCl3·xH2O was inactive and [{RuCl2(p-cymene)}] and catalysts, an effect that could also be achieved by applying
[{RuCl2(benzene)}2] were believed to be the relevant precursors. heat.[133] In 2018, Beller and co-workers[134] reported efficient
The reaction was promoted when PPh3 was added. A higher hydrogen generation from FA using RuH2(PPh3)4 without base
activity was achieved for RuBr3·H2O, with a TOF of 3630 h 1, additives (Figure 5). A TOF of 36000 h 1 was obtained by using
than for RuCl3·xH2O. However, deactivation of RuBr3·H2O C1 at 60 °C and the catalyst was stable for over 120 days. To

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2661 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Figure 4. Aromatic phosphines bearing sulfonate groups. Reaction conditions: RuCl3·xH2O (0.056 mmol), Ru/L = 1 : 2, HCOOH (9 mmol), HCOONa (1 mmol), H2O
(2.5 mL), 90 °C, 750 rpm.[126]

explore the cause of the high activity, Beller tested a series of


analogs of C1 (Figure 5). As C1 and C2 easily created vacant
coordination sites by elimination of triphenylphosphine or
hydrogen, they displayed better catalytic performances than
C3, C5, and C6 (Figure 5). The low activity of C4 (Figure 5)
demonstrated that Cl was not replaced, further indicating that
pre-activation by bases or other additives was not necessary.
Czaun et al.[135] tested a RuCl3 precursor with various
phosphine ligands for FA dehydrogenation in an emulsion of
FA, water, and organic solvents. The electronic effect of the
substituents had little effect on the reaction rate because similar
values were obtained for RuCl3 and PPh3, as well as for tris(4-
chlorophenyl)phosphine and tris(2-tolyl)phosphine. However,
tris(4-chlorophenyl)phosphine exhibited poor selectivity. The
solubility of the ligands influenced the decomposition rate, and
the high activity of PPh3·HBr likely resulted from the high
solubility of the ionic ligands. [Ru(HCOO2)2(CO)2(PPh3)2] [Ru-
(CO)3(PPh3)2] and [Ru(HCOO2)2(CO)4(PPh3)2] were isolated from
Figure 5. Ru complexes for FA dehydrogenation. Reaction conditions: THF the reaction system, but neither were comparable to in situ
(6 mL), catalyst (3.3 μmol), formic acid (6.6 mmol), H2O (13.2 mmol), catalysts in terms of both activity and selectivity. The effect of
T = 25 °C.[134]
surfactant addition was also investigated. Sodium dodecyl
sulfate improved the activity and selectivity of the catalysts. In
2016, Joó and Papp[136] synthesized cis-mer-[IrH2Cl(mtppm)3] for

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2662 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

selective aqueous FA decomposition to H2/CO2. The catalyst cantly influence solubility, the solubility decreases in the order:
displayed superior performance, achieving a high TOF of pTPPMS > mTPPMS, mTPPMS-Na3 > mTPPMS Na. In conclu-
298000 h 1 at 100 °C. sion, the activity of the catalytic system is always determined by
In addition to the aforementioned phenyl phosphine the synergistic effect of basicity, steric effects, and solubility of
ligands, some complex phosphines have been investigated. For the ligand.
most homogeneous FA dehydrogenation catalysts, bases are
necessary to initiate or accelerate the reaction. Reek and co-
workers[137] reported a new catalytic system in which bis- Catalysts with half-sandwich ligands
METAMORPhos ligands could act as an internal base. During
the study of the reaction mechanism, it was found that the Fukuzumi et al.[139] studied the selective decomposition of
ligands also promoted the coordination of FA and stabilization aqueous FA to H2 and CO2 at room temperature using a half-
of the transition states by hydrogen bonding. A TOF of 3092 h 1 sandwich Rh complex [RhIII(Cp*)(bpy)(H2O)]2 + (bpy = bipyridine,
was obtained at 85 °C in FA/toluene. he reaction rates differed Cp* = pentamethylcyclopentadienyl). Variations in the pH of the
significantly for different solvents because no external bases HCOOH/HCOONa solution in the range of 1–12 led to the
were added. Ligands were modified for further exploration.[138] highest TOF being recorded at pH = 3.8, corresponding to the
TOF values of 3090 h 1, 877 h 1, and 1791 h 1 were obtained for pKa of FA. A mechanism was proposed in which H2 was released
C7, C8, and C9, respectively (Figure 6). The activities were by the formation of [RhIII(Cp*){OC(O)H}(bpy)] +, followed by a
correlated with the electronic properties of the sulfonamide rate-determining β-hydrogen elimination step to yield [RhIII-
organic side groups. The reaction was facilitated by the electric (Cp*)(H)(bpy)] +. The comparable balance between the protic
donating group in the para-position (C7, nBu), but was and hydridic properties of [RhIII(Cp*)(H)(bpy)] + was believed to
hindered by the electron-withdrawing group in the para- be responsible for the efficient H2 generation observed. In 2012,
position (C8, CF3). the group synthesized a water-soluble [C,N]-cyclometalated
To date, phenyl phosphines have been widely studied and organoiridium complex, [IrIII(Cp*){4-(1H-pyrazol-1-yl-kN2)benzoic
new complex functional phosphines have attracted increasing acid-kC3}(H2O)]2SO4.[95] The reversible conversion of FA and H2/
attention owing to the elimination of external bases. For in situ- CO2 by the complex was achieved by changing the pH of the
generated catalysts, the ratio of ligands to precursors must be reaction mixture at ambient temperatures and pressures. A
accurate. Insufficient ligands for the catalytically active species maximum TOF of 1880 h 1 was obtained for the catalyst at
lead to poor stability. However, superfluous ligands reduce the 298 K and at a pH of 2.8. When the pH was exceeded 9, the
reaction rate. Some experiments have shown that molecule- catalyst was completely transformed into the corresponding
defined catalysts display better activity than the same catalytic benzoate complex, which displayed poor activity.
system generated in situ. These results indicate the importance During the research of transfer hydrogenation using FA as a
of the formation of active species. The basicity of the ligand has hydrogen donor, Himeda et al.[140] found that a water-soluble
a significant influence on its activity. Usually, ligands with half-sandwich catalyst [Cp*Rh(bpy)Cl]Cl promoted H2 evolution
higher basicity present a stronger σ-donating ability, promoting from FA at 40 °C with a TOF of 238 h 1. They further modified
dehydrogenation. The basicity can be modified by the sub- the bpy ligands and applied the half-sandwich catalyst (4,4’-
stitution of phosphines, but various steric effects may simulta- dihydroxy-2,2’-bipyridine and 4,7-dihydroxy-1,10-phenanthro-
neously be introduced. Some strong basic ligands (i. e., ligands line), thereby achieving efficient, waste-free, and recyclable
with nonaromatic cycloalkyl groups), are not as active as catalyst conversion from CO2 to formate.[141–143] The improved
expected due to poor solubility and steric effects. Most performance of the catalysts was attributed to the acid-base
phosphines have poor water solubility, implying that organic equilibrium of the hydroxy groups (Figure 7). Based on previous
solvents are required for the catalytic dehydrogenation of FA. work, they applied the 4,4’-dihydroxy-2,2’-bipyridine (DHBP)
Hence, the activity of the catalytic system is affected by ligand system to FA dehydrogenation.
solubility. Adding hydrophilic moieties to the ligands, such as Himeda and co-workers selected and modified a series of
sulfonate groups, can improve their solubility in water while N,N’-bidentate ligands. They significantly improved the activity
expanding the applicable scope of the phosphine. Because and stability of the catalysts while studying the mechanism in
both the position and number of hydrophilic groups signifi-

Figure 6. Iridium bis-METAMORPhos complexes for dehydrogenation of FA.


Reactions were carried out in toluene without base at 85 °C. TOFs were Figure 7. Acid–base equilibrium of 4,4’-dihydroxy-2,2’-bipyridine
determined between 4 and 30 % conversion.[137] complexes.[94]

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2663 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

detail. In 2009, Himeda and co-workers[94] reported a group of pendent base. Further theoretical and experimental studies
catalysts with different 4,4’-substituent-2,2’-bipyridine ligands demonstrated that the multifunctional ligands, which acted as
(Table 1, C10–C15). The highest TOF value of 14000 h 1 was proton-relays and strong π-donors, caused the high catalytic
obtained for C10 with DHBP, and no activity loss was observed activity. Wonglakhon and Surawatanawong[147] presented a
during five cycles. Owing to the acid-base equilibria of the mechanistic study of Cp*Ir complexes with proton-responsive
ligands, the reaction rate was affected by the pH of the FA/ ligands. Their research demonstrated the importance of alkali
HCOONa solution, and the reversible conversion of H2/CO2 to metal ions in the reversible CO2 conversion. Jaccob and co-
FA was achieved by controlling the pH. The reaction rates workers[148] proved the proton-responsive ligands were also
increased in the order Ru, Rh, and Ir using the same ligand. The compatible with Cp*Rh and Cp*Co. Himeda and co-workers[149]
effect of noble metal centers on the rate was consistent with conducted a detailed study of the reaction mechanism of a
previous reports on DHBP ligands in transfer catalytic system with bpy ligands. The kinetic isotope effect and
hydrogenation.[143,144] For Ir and Ru metal centers, better distribution of H2, HD, and D2 in the production gas were
catalytic performance of the system was observed using ligands compared using different combinations of deuterated FA and
with stronger electron-donating abilities. deuteroxide. The rate-determining step for the iridium complex
However, the activities appeared to be independent of the was H2 production from Ir hydrides while fast H-D exchange
substituents at the Rh metal center, which is likely due to the occurred via the hydrides. β-Hydrogen elimination was the rate-
different rate-determining steps involved.[145] In 2012, Hull, determining step for the Ru and Rh complexes. Later, Himeda
Fujita, Himeda, and co-workers[146] reported reversible intercon- and co-workers[150] synthesized Ir complexes with 6,6’-substitu-
version between CO2, FA and formate catalyzed by a homoge- ents-2,2’-bipridine ligands for FA dehydrogenation. The same
neous iridium catalyst with proton-responsive ligands at mild substituent effect as with 4,4’-substituents-2,2’-bipridine was
temperatures and pressures. C16 (Figure 8) was highly catalyti- observed, and the strong electron-donating substituents en-
cally active in aqueous FA dehydrogenation, achieving a TOF of hanced the reaction activity. However, the catalyst with
228000 h 1 at 90 °C and a TON of 308000 at 80 °C. The electron-donating -Me displayed a lower activity than the
monometallic complex C10 (Figure 8), with an activating group unsubstituted catalyst, suggesting that the significant steric
(hydroxy) but no pendent base, and the bimetallic complex C17 hindrance of methyl (Taft steric constant (Es) = 1.24[151])
(Figure 8), with neither an activating group nor pendent base, possibly impeded the coordination of FA. The less hindered
were compared with C16 to confirm the role of the ligands. MeO (Es = 0.55[151]) displayed no steric effect. The pH depend-
In the FA dehydrogenation reaction, C10 was much less ence of the catalysts also differed. Catalysts with OH groups in
active than C16, whereas C17 displayed similar activity to C16. the ortho position formed a proton relay, which assisted the
However, in the reversible CO2 hydrogenation reaction, C16 reaction of [Ir]-H by providing a proton, thereby lowering the
displayed an extraordinarily high activity compared to both C10 energy barrier of H2 evolution. Consequently, β-H elimination
and C17, illustrating the effect of the activating group and was the rate-determining step. The curve of the reaction rate
assumed a volcanic shape for these catalysts. However, for
catalysts with 4,4’-OH and non-responsive ligands, the rate-
Table 1. Complexes bearing 4,4’-substituted 2,2’-bipyridine ligands.[94] determining step was the release of H2, which was affected by
Catalyst structure Catalyst n M R TOF [h 1] the concentration of H +. Moreover, the reaction rate decreased
with increasing pH. Webster, Papish, and co-workers[152] also
studied the FA dehydrogenation using Ir and Ru complexes of
C10 5 Ir OH 2800[a]/2400[b]
C11 5 Rh OH 510[a]
N-heterocyclic carbene- and pyridinol-derived ligands. The
C12 6 Ru OH 40[a] important role played by dxbp-type (dxbp = dihydroxy-bipyr-
C13 5 Ir OMe 1200[b] idine or dimethoxy-bipyridine) ligands in the reaction was
C14 5 Ir Me 140[b]
C15 5 Ir H 30[b]
highlighted. To develop new, efficient catalysts for selective FA
decomposition, ligands with an increased electron-donating
[a] In 2 m aqueous FA solution (10 mL) at 60 °C; [b] In 1 m aqueous FA ability must be introduced. Accordingly, Wang, Himeda, co-
solution (10 mL) at 60 °C. workers[153] synthesized new Ir and Rh complex with five-
membered aromatic N-heterocycles (azoles) as ligands. The
replacement of pyridine in ligands by azoles evidently improved
the activity compared to that of C18 (Table 2).
The TOF of C20 was half that of C19 (Table 2), which was
attributed to the two adjacent N atoms in the pyrazole that
withdrew electrons into the aromatic ring, resulting in a lower
electron-donating ability of the coordinated N atoms compared
to that of imidazole. The electron-withdrawing effect of the
benzene ring caused the low activity of C21 (Table 2). The
reaction rate of C22 (Table 2) was higher than that of C18 and
even that of C10 with electron-donating -OH groups. The
Figure 8. Ir complex for aqueous FA dehydrogenation.[146] activity of C23 (Table 2) was further enhanced by introducing

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2664 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Table 2. [Cp*M(H2O)(L)]SO4 catalysts with different N,N’-bidentate ligand- containing pyridine-azole and pyrimidine-azole/azoline ligands.
s.[a][153] Highly robust H2 production from FA solution was accom-
Ligand structure Metal center Catalyst TOF[b] [h 1] plished. C26 was more active and more stable than C25
(Figure 9), confirming the key role that pendent OH played in
the reaction. The H2 release rate of C27 (Figure 9) was more
Ir C18 30[c] than double that of C26.
However, an evident decrease was observed for C28 (Fig-
ure 9) when changing the position of the N atom. The
substitution of pyrazole in C27 with imidazole in C29 (Figure 9)
Ir C19 810
led to decreased activity and decreased durability at a low pH
of 1.7, and C29 precipitated below a pH of 2.6. Interestingly,
C29 achieved a TOF of 19900 h 1 at a higher pH of 3.0. The TOF
Ir C20 400 of C29 was even higher than that of C27 (TOF of 18000 h 1 at
an optimized pH of 2.8). Although a TOF of 5520 h 1 at a pH of
1.7 was obtained for C30 (Figure 9) with pyridine-imidazoline
Ir C21 570 ligands, it increased to 32500 h 1 at a higher pH of 3. The
highest TOF of 322000 h 1 and TON of 2050000 were achieved
using C30 and C27, respectively. The experimental results
suggested that these ligands were highly active and stable and,
Ir C22 3980 therefore, showed potential for practical applications. The
deuterium kinetic effect and density functional theory (DFT)
calculations showed that the rate-determining step switched
between β-H elimination and H2 production when changing the
Ir C23 7020
Rh C24 250[d]
pH of the reaction solution. Efficient H2 generation from FA
using half-sandwich Ir complexes with 2,2’-bisimidazoline
[a] Reaction conditions: 1 m aqueous FA solution (10 mL), T = 60 °C. ligands bearing γ-NH groups was independently reported by
[b] Average value during the first 10 min. [c] Average value during the the groups of Himeda[155] and Li.[97] However, the role of the
first 2 h. [d] Used 2 m aqueous FA solution. amino groups acted during the reaction was not clearly studied.
As an extension of their previous study on biomimetic
complexes with pendant OH groups, Wang, Li, Bao and co-
four methyl substituents. A TOF of 34000 h 1 was obtained for workers designed and synthesized a series of Ir complexes with
C23 at 80 °C in aqueous FA in the absence of a base. Based on pendant N moieties and proposed an detailed mechanism.[156]
the success that catalysts containing bipyridine with pendent C31 (Figure 10) with pendant pyridine displayed better catalytic
OH groups and azole ligands achieved in the field of aqueous performance than C32 (Figure 10) containing phenyl, suggest-
FA dehydrogenation, Muckerman, Fujita, Himeda and co- ing the importance of the N atom of pendant pyridine. This
workers[154] designed a group of bio-inspired Ir catalysts effect was unusual, as the N atom was far away from the metal

Figure 9. Bio-inspired Ir complexes for aqueous FA dehydrogenation. Reaction conditions: 1 m FA solution (10 mL), catalyst (1 μmol), T = 60 °C. Average TOF
value during initial 10 min.[154]

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2665 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Figure 10. Complexes with pendent N-containing groups. Reaction conditions: 1 m aqueous FA solution, T = 60 °C. Average TOF value during initial 10 min.
* T = 80 °C.[156]

center. In addition, C33 (Figure 10) with methylated 1H-pyrazole pressures. The catalysts containing 4,4’-hydroxy-2,2’-bipyridine
substituents displayed a much lower TOF than those of C31 achieved a TOF of 3100 h 1 and TON of 100000 at 60 °C and
and C32, highlighting the importance of the N H proton. The atmospheric pressure.[94] Moreover, the TOF and TON reduced
reaction rate for C34 (Figure 10) with CF3 was faster than that to 2510 h 1 and 38100, respectively, at 80 °C and 30 MPa.
of unsubstituted [Cp*Ir(pz-py)(OH2)]2 + (pz-py: 2-(1H-pyrazol-3- Kawanami and co-workers[157] proposed that the deactivation
yl)pyridine). C35 (Figure 10) containing benzimidazole displayed was caused by catalysts precipitation owing to chelating
a TOF similar to that of C31. A TOF of 46510 h 1 was obtained conformation change of the bipyridine ligands. They presented
for C31 at 90 °C in a 1 m FA solution. The results showed that a Cp*Ir complex with 4,7-dihydroxy-1,10-phenanthroline, which
the synergistic effect of the N atom of pendent pyridine and prevented cis/trans isomerization of the pyridine skeleton by
N H moiety of pyrazole were not the only electron-withdrawing bridging. The catalytic activity was comparable to that of Cp*Ir
groups responsible for the high activity of C31. complexes containing bipyridine ligands with improved stabil-
A study of the mechanism demonstrated that the N H ity. A TON of 5000000 (2600 h) was achieved using 1 μmol of
moieties in the pyridinium cation and 1H-pyrazole unit formed catalyst in a 10 m aqueous FA solution at atmospheric pressure.
hydrogen bonds with FA molecules, which stabilized the Moreover, 8 μmol of the catalyst maintained its activity over 10
transition states and assisted proton transfer in the rate- cycles at a high pressure of 22 MPa. The catalyst could be
determining step. This is the first report of two functional N recycled easily because it precipitated after the consumption of
moieties promoting proton transfer in cooperation, which can FA due to its varying solubility with pH. Kawanami and co-
yield new ideas for designing catalysts. Himeda and co- workers[158] also investigated the kinetic and mechanism studies
workers[155] presented Cp*Ir complexes with bidentate ligands of Cp*Ir complexes containing DHBP for high-pressure H2 gas
including pyridyl-imidazoline and 2,2’-biimidazoline for efficient generation from FA solutions. They then developed a system
H2 evolution from FA. However, in large-scale experiments, a for continuous high-pressure H2 evolution and separation from
clear degradation of the catalyst with 2,2’-biimidazoline was FA at mild temperature.[61] In 2018, Himeda, Onishi, and co-
observed. Nevertheless, the catalyst with pyridyl-imidazoline workers[159] studied the selective decomposition of FA to H2/CO2
achieved a TON of 2000000 over 363 h. The results showed in water using picolinamide-based Ir catalysts. They revealed
that, although the pyridyl unit could not further enhance the the effect of the amide-N moiety on the activity and stability. In
reaction activity, it provided more durability than the imidazo- situ catalytic systems of the Ir precursors and L26–L34 (Table 3)
line unit. The FA dehydrogenation performance using catalysts were applied. The L26 system displayed a moderate activity.
containing N-methylated pyridyl-imidazoline ligands was almost The in situ catalytic system with an amidine moiety (L27)
the same as that using catalysts containing pristine ligands. DFT displayed an activity and pH dependence similar to that of C38
calculations supported the theory that the N H proton was not (Table 3). The results indicated that the amidine moiety and
involved in the reaction. For practical applications, catalysts for imidazoline served the same function in the reaction. The
FA dehydrogenation should remain active and durable at high

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2666 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Table 3. Catalysts and ligands for in situ catalytic systems tested in FA L28 and L29, specifically when sodium formates were added to
dehydrogenation.[159] the reaction solution. A high TOF of 118000 h 1 was obtained
Ligand/catalyst structure Ligand/catalyst Conversion TOF[a] [h 1] for the system with L28 in an 8 m FA/sodium formate (SF; 9 : 1)
solution at 60 °C. Systems containing L31–L34 were synthesized
to investigate the effect of five-membered N-heterocycles. The
low conversion ratio of L31, L33, and L34 (conversion of 52 in a
C37[b] 89 70400 1 m 1 : 1 FA solution) indicated that the catalysts degraded
during the reaction. Both the activity and stability were higher
with the N-phenyl ligand in L32 than in L31, which was
consistent with the results for L30 and L28. Because of the high
reactivities achieved using five-membered azole-type ring
ligands and pyridine-based ligands, catalysts with pendent OH
C38[b] 90 7230
ligands containing a hydroxy-pyrazole moiety were also tested
by the group.[160] Both the catalytic performance during CO2
hydrogenation and FA dehydrogenation were consistent with
the electron-donating abilities of the substituents of the
L26 90 2560 pyrazole moiety in the increasing order L35 < L36 < L37 < L38
(Table 4). Under similar conditions, catalysts with L38 displayed
higher activity than C10 and C19, indicating that OH had a
L27 90 10600 more beneficial effect on the pyrazole moiety than on the
pyridine moiety. In a continuous experiment, a TON of over 10
million during approximately 35 days was achieved using
L28 92 30200 catalysts containing L38, representing excellent durability.
Based on various N,N’-bidentate ligands, Kawanami and co-
workers[161] conducted research aiming to establish the correla-
tion between the stability and structure of ligands. L38–L44
L29 91 25300
(Figure 11) were chosen for the study.[94,97,159–161] The observed
activity trend revealed higher activities
for ligands with stronger electron-donating abilities. L38–
L30 92 61700 L42 were selected for further stability studies because of their

L31 61 5940 Table 4. Results of CO2 hydrogenation and FA dehydrogenation using


[Cp*Ir(H2O)(L)] with pyridyl-pyrazole derivatives.[160]
Ligand structure Ligand Cat.[a] TOF[a] Cat.[b] TOF[b]
[μmol] [h 1] [μmol] [h 1]

L32 92 34110
L35 200 20 500 240

L33 32 9090
L36 200 11 100 630

L34 91 12200

L37 20 130 100 2700

[a] Reaction was carried out using in situ or isolated catalysts (1 μmol) in a
1 m FA solution (10 mL) at 60 °C. Average TOF value during initial 5 min.
[b] Isolated catalyst was used.

L38 20 650 100 6720

complex containing L28 was efficient for both FA dehydrogen-


ation and CO2 hydrogenation.
[a] CO2 hydrogenation reaction conditions: PH2 =CO2 = 1.0 MPa (1 : 1), [NaH-
However, its catalytic performance was significantly affected CO3] = 1.0 m in aqueous solution at 50 °C. Average TOF value during first
by the substitution of N-amide. The catalytic system containing 1 h. [b] FA dehydrogenation reaction conditions: 1 m aqueous FA solution
L30 displayed higher activity than those of systems containing at 60 °C. Average TOF value during first 10 min.

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2667 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Figure 11. Selected ligands for exploration of relationship between ligands and stability. TOF values were measured at 60 °C and atmospheric pressure.[161]

high activities. The highest TOF was obtained for the Cp*Ir catalysts were divided into four groups: metal ions, cyclo-
complex containing L40, followed by L41, L39, L38, and L42. A metalated groups, donor groups, and R groups (the substitu-
different catalytic stability trend of (L39) > L42 > L38 > L40 > ents in the electron-donating N atom; Figure 12). The effect of
L41 was observed when batch experiments were conducted at the groups on the catalytic activity and the role they played
a high initial FA concentration and high pressure. The Ir during the reaction were investigated by substituting groups.
complexes containing L39 and L42 were the most stable, even Various donor groups were examined. The results indicated that
at 100 MPa.[61] The stability difference between L39 and L40 catalysts with imidamide-based donor groups containing an
indicated that the stability of the catalysts could be enhanced N H proton, i. e., 2-imidazolyl or 2-imidazolinyl, displayed high
by pyridine moieties, even at high pressures. The reaction activity; however, other N-heterocyclic donors (e. g., 2-oxazolinyl
solution with the catalyst containing L41 was analyzed by and 2-pyridyl), imine, or saturated benzylamine-
electrospray ionization mass spectrometry (ESI-MS). The pres- derived ligands displayed almost no activity. Replacing the
ence of [(Cp*Ir)2H3] + with a low activity was confirmed, which N H proton in the 2-imidazolinyl ligand with methyl, benzyl, or
likely deteriorated the performance. The generation of CO, formyl groups led to inactive catalysts. This result highlighted
confirmed by infrared (IR) spectroscopy of the reaction solution, the critical role of γ-NH in the reaction and showed that the
possibly caused the deactivation of catalysts containing L38.
However, specific mechanism was still under investigation. In
2020, Kawanami and co-workers[162] introduced amino and
alkylamino moieties into the bpy ligands. The generation of
high-pressure H2 from aqueous FA without base additives was
achieved using the amino-functionalized Cp*Ir complexes. The
activity was improved by 4,4’-diamino substituents. The
introduction of longer alkyl chains in the ligands (-NMe2, -NEt2)
further enhanced the activity, although the stability decreased.
Position had a different effect on the reaction activities of
amino and hydroxy groups. The decomposition rate of FA into
H2/CO2 was faster for Cp*Ir complexes with 6,6’-dihydroxy-2,2’-
bipyridine ligands than for the 4,4’-substituted analogs. How-
ever, the opposite result was obtained for the amino substitu-
ents. 1H NMR analysis revealed that protonation of ortho-amino
groups in bpy ligands had occurred to form -NH3 +, which
subsequently became an electron-withdrawing group, thereby
resulting in a reduced TOF.
In 2013, Xiao and co-workers[96] studied FA dehydrogenation Figure 12. Modular structure of Cp*Ir complexes proposed by Xiao and co-
in FA/triethylamine (TEA) using half-sandwich Ir complexes. The workers and the “optimized” structure.[96]

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2668 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

electron or steric effects did not cause poor activity. Electron-


donating groups substituted in the cyclometalated group
improved the reaction rate, whereas the activity was hindered
by electron-withdrawing groups. It was concluded that a higher
electron density around the metal ion improved the reaction
activity. Ru and Rh analogs were synthesized and tested but the
catalysts were unstable. A high TOF of 147000 h 1 was achieved
using the optimal catalyst at 40 °C. A study of the mechanism
revealed that unusual FA-assisted proton hopping occurred,
resulting in proton transfer from remote to proximal N atoms.
The protonation of the hydrides followed, accompanied by the
release of H2. Zhang and co-workers[163] also highlighted the
importance of γ-NH functional group by DFT calculation. In
summary, different parts of the ligands can be optimized
independently and then combined in a suitable structure by
Figure 13. Dioxime ligands screened in FA dehydrogenation. Reaction
the modular design of the catalysts presented here. This conditions: [Cp*IrCl2]2 (0.5 μmol), Ir/L = 1/1.2, 1 m FA solution (10 mL),
method is instructive for designing new catalysts. T = 60 °C. Average TOF value during initial 3 min. * Average TOF value during
In 2015, homogeneous Cp*Ir catalysts for base-free FA initial 20 min, owing to slow reaction rate.[164]
dehydrogenation with unprecedented high activity were re-
ported by Li and co-workers[97] A high TOF of 487500 h 1 was
obtained by using [Cp*Ir(2,2’-di-2-imidazoline)Cl]Cl at 90 °C, the sterically hindered coordination of FA due to the bulky
whereas an in situ catalytic system generated from [IrCp*Cl2]2 isopropyl groups. High activity was obtained for catalysts
and 2,2’-bis-1,4,5,6-tetrahydrophrimidine achieved a high TON containing L50–L52 (Figure 13). The catalyst containing L51
of 2400000 at 80 °C. Aromatic ligands, such as bipyridine, displayed the best catalytic performance without base addi-
bipyrimidine, and biimidazole, have been widely used as tives, and its stability was improved by an appropriate excess of
homogeneous half-sandwich catalysts in the field of aqueous ligands. However, for large-scale FA decomposition, the catalyst
FA dehydrogenation. The effectiveness of non-aromatic ligands containing L58 displayed the best performance. A high TON of
was demonstrated in this report, which extended the range of 3900000 and an average TOF of 65000 h 1 were achieved using
ligand types. Li, Xiao, and co-workers[99] also found that halide a catalyst containing L52 in a 10 m FA solution at 90 °C.
anions, particularly iodide anions, could promote H2 production Metal/NH bifunctional catalysts were proved to be effective
from FA/TEA using a [RhCp*Cl2]2 catalyst. This was attributed to in H2 transfer reaction, but it remained scarcely explored in the
the acceleration of the rate-determining step (hydride forma- field of FA dehydrogenation. Kayaki, Ikariya and co-workers[165]
tion) by I . Although significant progress has been achieved in synthesized amido and hydrido(amino) iridium complexes
the field of FA dehydrogenation, further improvement of the derived from N-triflyl-1,2-dephenylethylenediamine (TfDPEN).
catalyst stability is still required, specifically at the conditions The complexes decomposed FA to H2/CO2 selectively at mild
where the catalysts display their best performance (e. g., at high temperatures, even without base additives. The decreasing
temperatures). Protection and degasification, which are incon- activities in the order C39a > C39b > C39c > C39d revealed that
venient for practical application, are required for most reported electron-deficient substituents on the sulfonyl unit facilitated
catalysts. Li and co-workers[164] reported a new Ir catalyst with a the dehydrogenation reaction (Figure 14). C40, as the precursor
dioxime-derived ligand that exhibited high stability and of C39a, was active in the presence of a base, but no H2
efficiency for the additive-free dehydrogenation of FA. More- evolution was observed when using the Rh and Ru analogs.
over, all manipulations were conducted without protection and C43 (Figure 14) was confirmed to be an intermediate in the
degasification. In their previous work, catalysts containing N,N’- mechanistic study. The TOF of C44 (Figure 14) was significantly
diimine were proved to be effective for aqueous FA decom- lower than that of C43. The results demonstrated the important
position into H2/CO2,[97] but the complexes were unstable when role of N H protons in the reaction. In addition, proton-relay
the reaction temperature rising up. They proposed that the processes, mediated by N H protons and water, that promote
facile hydrolysis of the imine motif in cyclic structures under efficient H2 production have been proposed. Later, the groups
reaction conditions caused deactivation. To avoid this problem, of Kayaki and Ikariya investigated the deactivation mechanism
various dioxime ligands with diamine structures outside the of the catalysts. A higher TOF was obtained using C39a at
cycles were designed. The catalysts containing L45 (Figure 13) higher temperatures. However, a lower TON demonstrated that
exhibited a low initial TOF value. No activity was observed for thermal degradation may cause activity loss. The thermal
L46, containing electron-withdrawing chloride groups, whereas stability of C40, an important intermediate involved in the rate-
the [IrCp*Cl2]2 precursor with both electron-donating groups determining step, was investigated by heating in reflux dimeth-
substituted by L47 and L48 displayed enhanced activity yl ether (DME) for 38 h.[166] Two different iridacycle complexes,
(Figure 13). The results indicate that electron-rich substituents C45 and C46 (Figure 14), were successfully separated by silica
in glyoxime could enhance the reaction activity. The low activity gel column chromatography. Both C45 and C46 were less active
of catalysts containing L49 (Figure 13) was probably caused by than C39a, indicating a relationship between deactivation and

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2669 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Figure 14. Catalysts proposed by Kayaki, Ikariya and co-workers.[165,166]

cyclometalation. Thus, C47 (Figure 14), with no substituents for ment in activity and durability was observed for complexes
cyclometalation, was designed, and a remarkable increase in containing L56. A high TOF of 3290 h 1 was achieved at 90 °C
TON from 3190 for C39a at 25 °C to 15400 for C47 at 60 °C was for L57 (Table 5), which consisted of a cyclohexane ring with
achieved. In conclusion, their studies on the thermal behavior vicinal amines. The compounds containing L58 (Table 5)
of DPEN-based complexes indicated that the cyclometalation of displayed tempered activity and presented a similar steric
TfDPEN ligands might hinder H2 evolution from FA. This structure but altered electron properties compared to L57. A
information can promote the development of new bifunctional higher TOF was obtained with L58 than with L59 (Table 5),
catalysts with better activity and stability for a range of indicating that a larger bite angle of the ligand reduced the
asymmetric and non-asymmetric processes. activity. The electron density in the phenyl ring of L60 (Table 5)
Laurenczy and co-workers mainly studied FA dehydrogen- was reduced by the adjacent N atoms, resulting in a lower
ation using homogeneous catalysts with phosphine ligands. electron-donating ability that facilitated the redox reaction of
Because half-sandwich complexes were found to be effective complexes containing L60. Experiment results showed that the
for aqueous FA dehydrogenation, and based on the efficient, catalysts rapidly deactivated during the reaction. Iridium
base-free catalysts for H2 generation from FA in water reported compounds with more complex ligands, L61 and L62 (Table 5),
by Li and co-workers,[97] Laurenczy’s group began their work in displayed tempered activities. Although Rh complexes contain-
the field of half-sandwich catalysts for FA dehydrogenation. In ing L53, L57, and L58 were synthesized and tested, their
2017, Laurenczy and co-workers[167] published a study on H2 performance was not comparable to that of the Ir analogs in
production from aqueous FA solution using complexes contain- terms of both activity and stability. The main drawback of half-
ing different N,N’-donor ligands. Based on the role played by sandwich metal catalysts containing N,N’-donor ligands is their
the nitrogen atoms, the ligands were classified into three stability, particularly at high temperatures and pressures.
groups: aliphatic (α), conjugated (β), and aromatic (γ). For the Ir Moreover, the metal center is easily reduced to a black
metal precursor, moderate activity was achieved with the heterogeneous substrate, which was experimentally observed.
simplest ligands, L53 and L54 (Table 5), where the secondary Among Ru-arene complexes containing 1,3,5-triaza-7-phospha-
amine was used. Both the activity and stability reduced when triccyclo-[3.3.1.1] decane ligand (PTA), η6-areneruthenium(II)
using L55 (Table 5). L56 (Table 5) incorporated two phenyl (RAPTA) displayed catalytic properties for transfer
groups on the carbon atoms of the ethylene backbone that hydrogenation.[206] Fink, Laurenczy, and co-workers[168] explored
sterically blocked access to the coordinating sites of the ligands its activity during FA dehydrogenation. The selective decom-
and further enhanced the stability. Therefore, a clear improve- position of FA to H2 and CO2 was achieved using [(η6-

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2670 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Table 5. The N,N’-bidentate ligands screened by Laurenczy and co- Rh observed by NMR spectroscopy compared to Ir, multinuclear
workers.[167] NMR spectroscopy was employed to elucidate the mechanism.
Ligand structure Ligand Metal center TOF[a] [h 1] Ir complexes based on the picolinamide ancillary ligands
displayed high catalytic activity in various reactions, Macchioni
IrIII 1431 and co-workers[170] investigated their feasibility as catalysts for
L53α
RhIII –[b] H2 liberation from FA. TOF values of 29715 h 1 and 27590 h 1
were obtained for C48 and C49 (Figure 15) at 60 °C in a solution
L54α IrIII 1367 of FA and potassium formate, respectively. The activity order,
C49 � C48 @ C50, indicated that an increased electron density
L55α IrIII –[b]
at the metal center possibly improved the performance. Similar
results were obtained for C49 and C50, whereas C51 exhibited
poor stability (Figure 15).
Fischmeister and co-workers reported that Ir and Ru
L56α IrIII 1606 complexes containing dipyridylamine ligands were efficient
catalysts for the reduction of levulinic acid to γ-
valerolactone.[171] The hydrogen liberated during the rapid FA
dehydrogenation in the initial stage under base-free conditions
IrIII 3278
L57α was curtailed for the transfer hydrogenation of levulinic acid by
RhIII –[b]
FA. They also found that Ir complexes were more efficient
during transformation than the dipyridylamine-ruthenium
IrIII 1480 analogs.[172] Based on the results aforementioned mentioned,
L58β
RhIII –[b] Fischmeister and co-workers[173] more systematically investi-
gated the H2 production from aqueous FA under base-free
conditions using Ir complexes based on dipyridylamine ligands.
L59β IrIII 515 XRD analysis revealed that C56 (Figure 16) presented almost
the same bond lengths and bond angles as the parent,
dimethylamino-free C52 (Figure 16). Because no apparent steric
L60γ IrIII –[b] differences were observed, the variation in the electron proper-
ties of the ligands was assumed to be responsible for the
activity differences. The chloro complex, C52, was less active
L61γ IrIII 726 than the more electron-rich zwitterionic sulfato complex C53
(Figure 16). The same results were observed for C54/C55 and
C56/C57 (Figure 16) and were attributed to the low lability of
the chloride complexes. C53 displayed a significantly higher
L62γ IrIII 892
activity than C54, suggesting that the substituents of the
remote bridging amine and N H proton were probably
[a] Reaction conditions: medium pressure sapphire NMR tubes; deionized involved in the reaction mechanism. C57 (Figure 16) was
water (1.6 g), formic acid (350 mg), precatalysts (12 μmol), T = 90 °C. TOF
values measured for total decomposition of FA. [b] FA was not decom- chosen for further investigation, and it displayed the highest
posed completely. activity with a TOF of 12321 h 1 at a pH of 1.8, whereas the
optimized pH for compounds containing dipyridine and N,N’-
donor ligands was in the range of 3.5–4.[94,97,153] Zhou and co-
workers[174] designed a self-supporting fuel cell system in which
benzyldimethylamine)(Ru)(PTA)Cl2]Cl. The durability of the cata- H2 produced from FA was supplied to the fuel cell, and FA
lyst was improved by introducing a second PTA in situ. Rh dehydrogenation was powered by using the exhaust heat from
complexes [Cp*Rh(bis(pyrazol-1-yl)methane)Cl]Cl were applied the fuel cell. Highly active IrCp*Cl2bpym was applied to realize
in aqueous FA dehydrogenation by Fink and Laurenczy.[169] A the system, and a TOF of 7150 h 1 at 50 °C was obtained. The
TOF of 1086 h 1 was obtained. Owing to the higher activity of catalyst also exhibited excellent stability after several cycles.

Figure 15. Complexes based on picolinamide. Reaction conditions: [HCOOH] + [HCOOK] = 3 m, pH 3.0, T = 25 °C).[170]

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2671 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

reaction and that the unique transformation of the catalyst was


essential for catalytic reactivity.
Many Ir complexes were found to be effective for H2
production from FA. Ru complexes always present similar
properties to their Ir analogs. However, the price of Ru is more
competitive than that of Ir. Based on the progress achieved
using Ir complexes containing N,N’-donor ligands,[94,97] Huang
and co-workers[177] reported FA dehydrogenation using a newly
developed [Ru(p-cymene)(2,2’-biimidazoline)Cl]Cl catalyst. The
catalyst completely converted FA to H2/CO2 and displayed a
TOF of 12000 h 1 at 90 °C in a solution of FA and sodium
formate. Moreover, it achieved a TON of 350000 at optimized
conditions. A reaction mechanism involving two competitive
pathways for the key hydride transfer step was proposed based
on experimental results and DFT calculations. In 2018, Singh
and co-workers[178] reported half-sandwich Ru-arene complexes
based on N-donor chelating ligands for aqueous FA dehydro-
genation. The catalysts bearing 8-(N-methylamino)quinoline
(AmQ) C60–C62 displayed higher reactivity than C59 containing
an ethylenediamine ligand (Figure 17). Negligible difference
activity was observed between C60 and C61, indicating that the
arene rings played a negligible role in the reaction. The TOF of
C62 was four times higher than that of C60, whereas C63
(Figure 17) displayed poor activity, suggesting the importance
of protic ligands and availability of the N H moiety. The effect
of the N H proton was confirmed by the low reactivities of C64
and C65 (Figure 17), which contained no N-donor ligands.
Under optimized conditions, C62 achieved a TOF of 940 h 1 at
90 °C. The presence of [(η6-C6H6)Ru(k2-NpyNHMe-
2+ 6 2
MAmQ)(H2O)] and [(η -C6H6)Ru(k -NpyNMe-MAmQ)] + were
Figure 16. Efficient catalysts for base-free FA dehydrogenation. Reaction
conditions: FA (4 mmol), H2O (2 mL), catalyst (0.01 mol%), 80 °C, 10 min.[173] confirmed during the mechanistic study. These coordinatively
unsaturated compounds were identified as active species
during catalysis, and the dimer form of the intermediate,
possibly a catalytic resting state, was isolated and successfully
The produced gas could be directly applied in the proton- characterized.
exchange membrane fuel cell, and an energy density compara- The ligands’ effect on reactivity is presented through the
ble to that of pure hydrogen was achieved. examples shown in Figure 18. Substitution affects the activity
In addition to half-sandwich catalysts featuring Cp*, other through electron donation, and the catalytic performance is
half-sandwich ligands have also been explored as catalysts for always consistent with the electron-donating ability of the
the selective decomposition of FA to H2/CO2. Williams and co- substituent. However, in some cases, steric hindrance should
workers[175] synthesized a half-sandwich iridium catalyst contain- also be considered. Some substituents play a more important
ing 2,2’-[(di-tert-butylphosphino)methyl]pyridine with [(COD) role in the reaction, such as pendent OH and N moieties. They
IrCl]2 as a precursor (COD = 1,5-cyclooctadiene). The catalyst can form unique intermediates that lower the reaction barrier
was stable in air and remained active even when exposed to air to further promote the reaction. The replacement of pyridine by
for two weeks. Under suboptimal conditions, low catalyst coordinated N atoms with a high electron-donating ability is
loadings (0.13 μmol) resulted in a TON of 2160000 during 40 another method to secure ligands with a strong electron-
cycles over four months, and a maximum TOF of 13320 h 1 was donating ability. Several five-membered N-heterocyclic rings
obtained. After the catalysis of neat FA to H2 and CO2, the and amide-N in picolinamide were found to be effective. The
catalyst precipitated once the consumption of the FA was combination of the aforementioned two methods can lead to
complete, which is convenient for recycling. However, CO was the design of new ligands. The species resulting from the
detected in the produced gas. The production of CO was various combinations may even show unexpected reactivity. A
inhibited by lowering the reaction temperature and adding a change in the chelating conformation of the ligand skeleton
base and water. In a study regarding the mechanism, the loss of may result in deactivation. Therefore, a ligand that prevents the
the cyclooctadiene ligand of the catalyst as cyclooctene was cis/trans isomerization of the pyridine skeleton by bridging was
identified and dimerization occurred. A more detailed inves- proposed. In addition, such ligands ensure high durability.
tigation was conducted by Williams and co-workers[176] They Half-sandwich metal catalysts have been successfully devel-
found that C2-twisted Ir dimers were formed during the oped. Many of these catalysts are highly active in aqueous FA

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2672 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Figure 17. Half-sandwich Ru-arene complexes (Reaction conditions: 0.01 mmol catalyst, 0.4 M, 2.5 mL FA, 0.005 mol HCOONa, at 90 °C. [a] 0.05 mmol
catalyst)[178]

Figure 18. The evolution of some ligands for aqueous FA dehydrogenation.[94,146,153,154,156,157,159,160]

dehydrogenation and give TOF values of over 105 h 1 under can be built. Volatile organic solvents/bases, such as TEA, which
mild conditions (< 100 °C). As most of the catalysts are water are harmful to the proton-exchange membrane of hydrogen
soluble, a catalytic system based on aqueous FA and formate fuel cells, should be avoided. The N,N’-donor ligands that are

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2673 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

most studied include bipyridine, biimidazole, and diamine. The several catalysts (Figure 19), including in situ catalysts gener-
increased electron-donating ability of these ligands was found ated from Ru(acac)3 precursors and one equivalent ligands,
to be vital for high catalytic reactivity. A high electron density molecular catalysts [Ru(k3-triphos)(MeCN)3](OTf)2 (C66), [Ru(k4-
around the metal center can also protect metal ions against NP3)(Cl)2] (C67), and [Ru(k3-triphos)(CO)(H)2] (C68). No gas
reduction, which can improve the stability of the catalyst. evolution was observed for the in situ generated catalysts at
However, insufficient stability remains a major obstacle for the 40 °C in the FA/TEA mixture, indicating that catalytically active
practical application of these catalysts. Some researchers have species were not formed from the precursors and ligands under
noticed this deactivation problem and have proposed hypoth- mild conditions. The highest activity was obtained for C66,
eses and designed catalysts with high durability. However, the which was a Ru complex containing strongly coordinating
specific mechanism and solution to catalyst deactivation require tripodal polyphosphine and three labile solvent ligands. All
further investigation. catalysts generated in situ or in the molecular form displayed
decreased reactivity compared to C66. The catalytic perform-
ance of C66 was attributed to the availability of up to three
Catalysts with chelating ligands coordination sites for substrate coordination and activation.
Further mechanistic studies revealed that all metal-hydrido
To build an energy storage system based on FA, the FA species, possibly generated from triphos-based catalysts, were
dehydrogenation reaction and CO2 hydrogenation reaction rapidly converted on an NMR timescale. However, a stable
must proceed at comparable rates. In 2011, Morokuma, Nozaki, metal-hydrido was observed when using catalysts based on NP3
and co-workers[179] studied the reversible conversion of CO2 and ligands, where two competitive pathways involving the inter-
FA catalyzed by a PNP-ligated iridium(III) trihydride complex. mediate existed. The coexistence of the two competitive
The catalyst could selectively decompose FA, displaying an pathways likely degraded the catalytic performance. A more
optimized TOF of 12000 h 1 at 80 °C in tBuOH with TEA. A high detailed mechanism was presented by Perizzoni, Gonsalvi,
activity was also obtained for CO2 hydrogenation. The mecha- Beller and co-workers[181] They found that the amounts and
nism proposed, based on DFT calculations and NMR results, properties of the different phosphine ligands (triphos and NP3)
provided an appropriate explanation of the effects of hydrogen and ancillary ligands (Cl and MeCN) led to differences in the
pressure, base additives, and solvents on the activity. available coordination sites. DFT calculations showed that the
Many catalytic systems based on various Ru precursors and reaction proceeded via metal-centered (inner-sphere) or ligand-
diphosphines were successfully applied in the field of H2 centered (outer-sphere) pathways, depending on the ligands
generation from FA.[67,131] The tetradentate ligand tris-[2-(diphe- present. Iron(II) complexes containing linear tetraphosphine
nylphosphino)ethyl]phosphine (PPh3) stabilized Fe-based cata- ligands (P4 where P4 = 1,1,4,7,10,10-hexaphenyl-1,4,7,10-tetra-
lysts and efficient FA dehydrogenation was achieved using non- phosphadecane) for the selective decomposition of FA into H2/
noble metal catalysts.[73] Based on the research above, Gonsalvi CO2 were reported by Gonsalvi and co-workers,[80] but the
and co-workers explored the feasibility of Ru complexes catalysts were unstable. Based on the study by Beller and co-
containing polydentate tripodal ligands 1,1,1-tris- workers,[182] Gonsalvi and co-workers[183] combined P4 ligands
(diphenylphosphinomethyl)ethane (triphos; L63) and tris-[2- with Ru precursors to develop more durable catalysts for long-
(diphenylphosphino)ethyl]amine (NP3; L64) for FA dehydrogen- term FA dehydrogenation. Rac-P4 and meso-P4 isomers of linear
ation in a mixture of FA and organic amine (Figure 19). tetraphosphines were employed as ligands. In situ catalysts of
Peruzzini, Gonsalvi, and co-workers[180] studied the activities of [RuCl2(benzene)]2 containing meso-P4 displayed significantly
higher activities than catalysts containing rac-P4. The induction
times of all the catalytic systems were determined. A TON of
220848 was obtained for the in situ [RuCl2(benzene)]2/meso-P4
system in a continuous experiment. The well-defined Ru
complexes had shorter induction times than their in situ
analogs, resulting in higher initial TOFs. The mechanistic study
highlighted the active intermediate, trans-[Ru(H)2(meso-P4)] and
revealed that HCOOH activation occurred in only one H atom of
the octahedral complex because of the trans effect of the
second hydrido ligand.
Laurenczy, Li, and co-workers,[184] also conducted research
on catalysts with Ru precursors containing triphos. They
speculated that Lewis acids might have an unexpected effect
on the catalysts. Experimental results showed that the TOF of
Ru-triphos could be improved from 550 h 1 to 1920 h 1 at 90 °C
using Al(OTf)3. Moreover, no activity loss was observed by the
Figure 19. Ru complexes bearing tripodal ligands triphos and NP3. Reaction
conditions: HCOOH/OctNMe2 (11 : 10, 2.73 mL), preformed catalyst 14th cycle. The promotion of the activity and stability by Al(OTf)3
(12.9 μmol), T = 80 °C. Average TOF value after 1 h. *Precursor Ru(acac)3, was attributed to its ability to activate pre-catalysts and FA and
(12.9 μmol), ligand (12.9 μmol).[180] to stabilize the active Ru ion species. Based on the above

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2674 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Figure 20. Rh complexes for FA dehydrogenation.[188]

progress achieved using Ru precursors and triphosphine ligands release step. A very low conversion was obtained for the
in the field of FA dehydrogenation, Araugo and co-workers[185] complex [Rh(Cl)(CO)(PNH)] (C70, Figure 20) containing a biden-
recently presented triphosphine-containing Ru-acetato com- tate PNH ligand (PNH = 2-2-bis(1,1-dimethylethyl)
plexes. [RuCl(k2-OAc)(triphos)] (triphos = 1,1,1-tris phosphinomethyl-6-methylpyridine), which lacked the flanking
(diphenylphosphinomethyl) ethane)(1) and [RuCl(k2-OAc)(etp)] phenyl arm. The blocking of the fourth coordination site
(etp = bis[2-(diphenylphosphino)ethyl]phenylphosphine)(2) adjacent to the chloride ligands likely deteriorated the activity.
were fully characterized. Cyclic voltammetry results and struc- The known Rh complexes [Rh(CO)(PNN*)] [PNN* = 6-di(tert-
tural analysis indicated a higher electron density of (1), which butyl)phosphinomethine-2,2’-bipyridine; C71, Figure 20] and
was consistent with the higher reactivity of (1) compared to [Rh(CO)(PCP)] [PCP = 2,6-bis(di-tert-butylphosphinomethyl)
that of (2). In addition, the highest-occupied-molecular-orbital pyridine; C72, Figure 20] hardly displayed any activity, indicat-
vs lowest-occupied-molecular-orbital (HOMO-LUMO) gap of (2) ing that low-coordinate geometries and ligands with suitable
may also contribute to its poor activity. The effect of the solvent denticity were crucial.
was then investigated. H2 evolution proceeded faster in Most homogeneous catalysts studied for the selective
tetrahydrofuran (THF) than in MeOH or iPrOH. This result was decomposition of FA into H2/CO2 are air- or water-sensitive.
expected owing to the stronger coordination abilities of MeOH Hence, water-free and oxygen-free conditions are required for
and iPrOH competing with formate for the ruthenium center catalytic FA dehydrogenation. However, common FA typically
than those of THF. A TOF of 833 h 1 was obtained for (1) at contains a certain amount of water, implying that air- and
80 °C in THF. moisture-stable catalysts are more practical. Zheng, Huang, and
Except for energy generation, the missing linkage in a co-workers[189] reported a new class of Ru complexes containing
sustainable energy chain is the storage and liberation of energy. a pyridine-based ligand which could be dearomatized through
Plietker and co-workers[186] designed a rechargeable hydrogen the deprotonation of one NH-PtBu2 arm. C73–C75 (Figure 21)
battery. Charging and discharging were achieved through exhibited appropriate stability in water and oxygen.
reversible conversion between H2 and FA. The reverse process A TOF exceeding 7000 h 1 over a period of 9000 min and a
was catalyzed by the same (PNNP)(acetonitrile)RuII complexes. TON of more than 1 million at 90 °C were obtained with the
The catalyst was initially used in the selective oxidation of best-performing catalyst (C75). The solvent affected the reac-
C H.[187] It was then speculated that this structure could be tivity and lifetime of the catalysts, and better results were
active for hydrogenation. In the CO2 hydrogenation experiment, obtained in DMSO than in toluene. The reaction rate was
0.0015 mol% of catalyst reduced CO2 corresponding to accelerated by the addition of an external base, but this led to a
65.7 mmol 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) formate decrease in the lifetime. The results suggested that stability
within 4 h at 100 °C. The decomposition of FA was completed in may be influenced by the acid/base balance. Moreover, NMR
70 min at 100 °C under pressure-free conditions, and no CO was results indicated that the N H bond was involved in modulat-
detected. The reaction system displayed great durability, the
storage capacity did not drop over five charging-discharging
cycles, and the resulting H2/CO2 mixture could be directly
employed in fuel cells.
Reversible cyclometalation is considered as a strategy for
the activation of protic small molecules. Vlugt and co-
workers[188] introduced this strategy to the field of FA
dehydrogenation. Several Rh complexes containing reactive
cyclometalated ligands were synthesized and tested. [Rh(CO)L]
[L = 2-di(tert-butylphosphinomethyl)-6-phenylpyridine; C69,
Figure 20] attained a TOF of 169 h 1 at 75 °C in FA/dioxane,
and the external base had a negligible effect on the Figure 21. A new class of PN3-Ru complexes for H2 generation from FA.
decomposition rate of FA. DFT calculations demonstrated that Reaction conditions: Ru complex (1.0 μmol) DMSO (5.0 mL), 95 % commercial
the cyclometalated complexes were highly active in the H2 FA (injected by microinjector), T = 50 °C.[189]

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2675 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

Table 6. Bifunctional IrIII-PC(sp3)P catalysts for H2 generation from FA.[190]


Ligand structure Catalyst FG TOF[a] [h 1] TON[b]

C76 OH 13710[c]/9460[d] 318500[c]/189500[d]


C77 NH2 1200[c]/18390[d] 35080[c]/498190[d]
C78 COOH ND[c]/ND[d] 1452[c]/560[d]

[a] Initial TOF (average of 2 runs). [b] Maximal TON. [c] FA/TEA 5 : 3 in DME at 70 °C under air. [d] FA/SF 1:0.3 in DME at 70 °C under air.

ing the acidity of the reaction environment. Further mechanistic substituents can prevent the decomposition of catalysts and
studies revealed that two equivalent water molecules were formation of metallic byproducts. 2) The strong σ-donating
required for the activation of the catalysts and that the ability of the ligand can lead to an electron-rich metal center,
oxidation state of Ru did not change during the reaction. The which is beneficial for the dehydrogenation reaction.[54] Based
extraordinary water and air stabilities were determined by the on the report on Ir-NHC complexes for catalytic H2 production
tridentate binding of the ligand, steady + 2 oxidation state of from formate,[57] as well as their previous work on the Ir IPr
RuII, steric effect of the bulky tert-butyl moieties, and modu- (IPr = 1,3-bis(2,6-diisopropylphenyl)imidazole-2-ylidene)[191] scaf-
lation of the local environment through the two N H arms. fold, Iglesias, Oro, and co-workers[54] explored the use of Ir IPr
Efficient catalytic FA dehydrogenation without additives may complexes in FA dehydrogenation. C79 and C80 featured a
lead to a higher hydrogen capacity and more convenient waste labile CH3CN ligand trans to the hydrides, whereas the same
disposal. Although progress has been made in the field of H2 coordination sites were occupied by a chelate ligand (py) in
generation from FA under additive-free conditions,[167,177] the C81 and C82 (Figure 22). In addition, the positions trans to NHC
TON and TOF of additive-free catalysts still require improve- in C81 and C82 were occupied by a stable phosphine ligand
ment, and a breakthrough involving new catalysts operating by and a labile ligand (pyridine), respectively. C83–C85 (Figure 22)
less traditional mechanistic pathways is necessary.[138] In 2017, were designed to evaluate the potential effect of the hydrogen
Schapiro, Gelman, and co-workers[190] reported an efficient Ir bond interactions between the ligands and substrate or solvent.
catalyst system for FA dehydrogenation without additives. The The highest activity was obtained for C81 with a TOF of
complexes featured a modular dibenzobarrelene-based pincer 3870 h 1 at 80 °C in DMF. The high activity was attributed to the
ligand possessing a pendent functional group [PC(sp3)P]. The double coordination sites resulting from the two labile ligands.
effects of the different substituents in the outer coordination As C79 was unstable owing to the absence of strongly
sphere of the ligands on the activity were investigated using a
neutral OH-containing catalyst (C76, Table 6), a basic NH2-
containing catalyst (C77, Table 6), and an acidic CO2H-contain-
ing catalyst (C78, Table 6). C78 displayed the lowest activity
under all tested conditions and different reaction mixtures. C76
achieved a high TOF of 13710 h 1 and TON of 318000 at 70 °C
in 5 m FA/TEA (5 : 3) with 1,2-dimethoxyethane. The TOF
obtained for C77 was lower by an order of magnitude.
However, a contrasting result was obtained when the reaction
proceeded without amine additives. The TOF and TON of C77
were both double those of C75 in the FA/SF reaction system.
C77 decomposed neat FA with a TOF of 11760 h 1 and TON of
383000 at 70 °C, whereas no gas evolution was observed for
C75 under the same conditions. The mechanistic study
demonstrated that hydrogen liberation proceeded through
intramolecular protonolysis of the Ir hydride species using
remote functionality. The regeneration of active intermediates
proceeded by the nonclassical, outer-sphere, intramolecular β-H
elimination of CO2.
Ir complexes containing NHC (NHC = N-heterocyclic car-
bene) chelate ligands have been successful in many domains.
NHC ligands present ideal properties for the development of Figure 22. Ir complexes proposed by Iglesias, Oro, and co-workers. Reaction
homogeneous catalysts for FA dehydrogenation. 1) Strong M C conditions: FA (0.5 mmol), Ir catalyst (0.5 mol%), HCOONa (5 mol%), H2O
bonds and steric protection enhanced by the fan-shaped N- (1 mL), T = 80 °C. Average TOF value for first 1 min. N/R = no reaction.[54,192]

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2676 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

coordinated ligands, it displayed poor activity. C82 with no (AMP)(dppb)Cl2] and [Ru(AMP)(PPh3)2Cl2]; AMP = 4-(aminometh-
labile ligands had difficulty forming coordination sites, exhibit- yl)pyridine), were also tested, but all displayed a significant
ing the lowest activity. The higher reactivity of C81 compared decrease in activity, indicating the importance of the cyclo-
to that of C82 was explained by the possible generation of metalated CNN ligands. Complexes with CO ligands, i. e.,
coordination sites from the labile py trans position to NHC. [Ru(AMP)(CO)2Cl2], [Ru(k2-CNN)(CO)2Cl], and [Ru(CNN)(CO)2Cl],
However, C83–C85, with the same coordination position, were also synthesized, but displayed unsatisfactory TONs. In
presented average activities. C83 displayed the highest activity addition, CO was detected in the gaseous product and may
when the reaction proceeded in water, suggesting a different have resulted from the cleavage of the CO ligands from the
mechanism from that in DMF. Moreover, proton reduction may metal or dehydration reaction being favored. Recently, Milstein
occur in water. The replacement of the NH2 group by N(CH3)2 in and co-workers[198] achieved a major breakthrough in the field
C83 resulted in a sharp decrease in activity. The results of neat FA dehydrogenation. They synthesized a Ru 9H-acridine
highlight the importance of easily accessible coordination sites complex that exhibited high activity, stability, and selectivity.
and protic ligands. Iglesias, Oro, and co-workers[192] also C90 (Figure 23) maintained its activity during an intermittent
designed Ir catalysts for the solvent-free dehydrogenation of FA reaction period of two months, achieving a TON of 1701150. In
to avoid the overall hydrogen decline of the storage system addition, it presented appropriate tolerance for high pressure
caused by the solvent. The [Ir(PCP)(COD)]BF4 (PCP = 1,3-bis(2- and water. When the precipitates were analyzed by NMR
(diphenylphosphanyl)ethyl)-2-methyleneimidazoline) complex spectroscopy after the consumption of the FA, C90 was found
featured a PCP ligand based on an N-heterocyclic olefin (NHO) to be the major species. A biscarbonyl 9H-acridine complex C91
scaffold. C86 (Figure 22) proved to be the most effective (Figure 23), which was possibly formed through coordination of
catalyst in the absence of a solvent. The performance of C87 the trace CO generated, was also formed (below the detection
(Figure 22) was clearly worse than that of C86, possibly due to limit of the gas chromatograph). C91 also decomposed FA into
coordination site blockage by the strongly coordinated CO H2/CO2, but at a lower rate. Two pathways were proposed for
groups. Both C88 and C89 (Figure 22) did not contain NHO C90 and C91 because of the number of coordination sites
moieties and displayed lower activities than their NHO-contain- available. The aromatic hydrido-chloride complex C92 (Fig-
ing analogs. Further study revealed an enhancement in the ure 23) was also active during neat FA dehydrogenation and
activity of C86 in the presence of water. may react via a similar pathway to C91, albeit at a lower rate.
Ru complexes based on PNP chelate ligands have been Catalysts with chelating ligands have the most species and
effectively applied in many dehydrogenation and hydrogena- the most complex structure, which makes it difficult to
tion reactions in the past decade. The formation of active imido determine the specific relationship between structure and
or dearomatized intermediates is vital for catalytic (de) activity. However, the high variability of the structure is brought
hydrogenation. However, complexes with N-alkylated aliphatic in at the same time. Breakthroughs have been made in the field
pincer ligands were less likely to complete the process. of non-noble metal catalysts for FA hydrogenation and in base/
Interestingly, Czaun, Prakash, Olah, and co-workers[193] pre- solvent-free H2 generation from FA. Many reactions proceed
sented a pincer complex for formate dehydrogenation, the through untraditional mechanisms in which both the metal
activity of which was independent of the N H moiety. Beller centers and ligands are involved in the reaction. This may
and co-workers[194] found that, in aqueous methanol reforming, change or promote the rate-determining step. The number of
Ru complexes containing N-methylated ligands displayed lower available coordination sites is important to the activity. Some
activity than those containing free N H moieties. However, the strongly coordinated ligands may fail to form the free
opposite tendency was observed in FA dehydrogenation. coordination sites, and present a low activity. However, they
Consequently, Beller and co-workers[195] compared the function could contribute to the stability. Hence, the appropriate
of Ru PNP complexes containing N-methylated and NH ligands coordinating ability of the ligands is important.
in FA dehydrogenation. Catalysts with N-Me ligands exhibited
better performance under all pH conditions, and a TOF of
6492 h 1 was achieved at 92 °C. The higher TOF of Ru Other catalysts
complexes with N-Me ligands was also consistent with a lower
reaction activation energy. However, similar behavior was not Wills and co-workers,[199–201] using FA/TEA as a hydrogen source,
observed for the Fe and Mn analogs. [Ru(k3-CNN)(dppb)Cl] was found that Ru- and Rh-tethered catalysts could generate a gas
reported by Baratta and co-workers[196] as an efficient catalysts
for transfer hydrogenation reactions. Because of the interest in
the potential of cyclometalated Ru complexes for FA dehydro-
genation and the design of new catalysts, Beller and co-
workers[197] explored the complex in the field of FA dehydrogen-
ation. [Ru(k3-CNN)(dppb)(OOCH)] was determined to be an
active intermediate with a TON of 9085 over 3 h, which is
higher than that with [Ru(k3-CNN)(dppb)Cl]. Several analogs
without CNN ligands ([Ru(PPh3)Cl2] and [Ru(PPh3)(dppb)Cl]), as
well as complexes with AMP ligands instead of CNN ([Ru- Figure 23. Ru complexes for neat FA dehydrogenation.[198]

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2677 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

mixture of 1 : 1 H2 and CO2 at room temperature by transfer present investigation suggests that ligands that increase the
hydrogenation. Other catalysts were explored in the FA-TEA electron density of the metal center can promote both activity
system (FA: TEA = 5 : 2). However, deactivation occurred before and stability. The electron-donating substitution of the P atom
the complete consumption of FA, indicating the unsuitability of can increase the basicity of ligands to further enhance activity.
the catalysts. Wills and co-workers[202] speculated that Ru and Rh However, steric effects and solubility simultaneously play a
complexes with a heteroatom adjacent to the metal are crucial role. The same applies to the use of N,N’-bidentate
possible candidates for catalyzing H2 production in the FA/TEA ligands. The direct application of a ligand structure with a
system. Various catalysts were screened. A TOF of 18000 h 1 higher electron density is another approach. Substitution can
was obtained for RuCl2(DMSO)4 at 120 °C, and the activity was also change the rate-determining step. Some pendent substitu-
maintained for four cycles. RuCl3·H2O and [RuCl2(NH3)6] achieved ents interact with FA or water molecules, assisting in the
TOFs of 17400 h 1 and 18000 h 1, respectively. Investigation of reaction through creation of a lower-energy pathway. Bio-
the mechanism revealed that the same pre-catalyst inspired catalysts and those that react via nonclassical mecha-
[Ru2(HCO2)4(CO)4], a bis-ruthenium complex containing bridging nisms in which both the ligands and metal centers are involved
formate ligands and carbonyl ligands, was formed for all in the reaction display unexpected reactivity, specifically for
precursors. However, temperatures exceeding 100 °C were additive-free FA dehydrogenation. Many known catalysts for
required for the formation of the pre-catalyst, which resulted in transfer hydrogenation are potential candidates for this applica-
the simultaneous production of CO. Consequently, monomeric tion.
Ru complexes with lower activity appeared to have been Over the past decade, tremendous progress has been made
formed. As FA decomposed into H2 and CO2, the boiling point by using homogeneous noble metal catalysts. Researchers have
of the FA/TEA mixture decreased, resulting in the loss of TEA by designed and synthesized complexes containing various ligands
evaporation at higher temperatures. Kendall, Wills, and co- to achieve the goals for practical application of FA as a
workers[203] sought to find an alternative with a higher boiling hydrogen carrier, which include high activity under mild
point to retain the integrity of the reaction mixture during long- conditions, high selectivity for FA decomposition into H2/CO2,
term continuous H2 evolution. They presented a mixture of FA high hydrogen content (additive-free), and low cost.
and a tertiary amine base to realize efficient, continuous The optimal conditions to achieve the highest activities of
generation of H2/CO2. They applied either a temperature- or many catalysts include the addition of solvents or exogenous
impedance-based feedback system to control the FA charging bases that lower the hydrogen content. However, when
rate. additives with negligible consumption are added to the reactor
Owing to the significant influence of non-classical dihydro- and the storage tank is loaded with neat FA with a high
gen bonds on the reactivity and stereochemistry of organo- hydrogen content, the overall capacity does not decrease
metallic hydride compounds, as well as the unique chemical significantly. Notably, volatile solvent and organic matter must
properties of bimetallic complexes, Lau and co-workers[204] be removed for fuel cell development to be feasible. Research-
synthesized and characterized dihydrogen-bonded heterobime- ers have developed many catalysts with non-noble metal
tallic complexes [(η5-C5R5)Ru(CO)(μ-dppm)M(η5-C5R5)](M = Mo, centers due to the high cost of noble metals. They have
W; R = H, CH3). The dehydrogenation of FA was investigated achieved great success, but improvements in both activity and
with these complexes. The poor activity could was attributed stability are still required. Although high cost can be a problem
the formation of a stable formate intermediate. when considering a complete catalyst recovery system, the
In 2010, Fukuzumi and co-workers[205] revealed a hetero- price is not the major obstacle for practical application. A TOF
dinuclear Ir Ru complex, [IrIII(Cp*)(H2O)(2,2’-bipyrimidine) of over 105 can easily be achieved by using efficient,
RuII(2,2’-bioyridine)2](SO4)2, for aqueous FA dehydrogenation. A homogeneous catalysts with noble metal centers at 60 °C and
TOF of 426 h 1 was achieved for the catalyst at 298 K. An without CO detection in the gas produced, indicating viable
unusually strong tunneling effect was observed in kinetic and activity and selectivity under mild conditions. However, chal-
mechanistic studies. This was the first report of a tunneling lenges relating to the stability of the catalysts remain.
effect in the field of homogeneous catalytic FA dehydrogen- Deterioration was observed in many catalysts with excellent
ation. catalytic performances after several cycles under laboratory
conditions, indicating the lability of the catalytically active
species. Catalysts that can achieve continuous H2 evolution
Conclusion and Outlook from FA over a long period usually exhibit lower activities than
at optimized operating conditions. Moreover, the durability of
In this Review, we have focused on the structure–function catalysts under resting conditions should also be considered as
relationships between ligands and catalyst activities. We aimed the catalysts remain in the reaction mixture continuously,
to provide an overview of the effects of ligand perturbations on thereby experiencing entire cycles of working and resting.
catalyst activity. In addition, the application potential of main- However, this issue has scarcely been investigated. In addition,
stream H2 storage methods, particularly that of FA as an on- most catalysts exhibit instability towards oxygen and moisture.
board hydrogen carrier, was explored. Some researchers have reported this problem and have
Researchers have designed several homogeneous catalysts proposed deactivation mechanisms and effective solutions.
and have studied their structure–function relationships. The New catalysts with high TONs have recently been reported.

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2678 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

However, none of the catalysts appear to meet the practical [34] M. Hirscher, V. A. Yartys, M. Baricco, J. Bellosta Von Colbe, D. Blanchard,
R. C. Bowman, D. P. Broom, C. E. Buckley, F. Chang, P. Chen, Y. W. Cho,
requirements. Improvements are still required to enhance the
J. Crivello, F. Cuevas, W. I. F. David, P. E. de Jongh, R. V. Denys, M.
stability of the catalysts. Dornheim, M. Felderhoff, Y. Filinchuk, G. E. Froudakis, D. M. Grant, E. M.
Gray, B. C. Hauback, T. He, T. D. Humphries, T. R. Jensen, S. Kim, Y.
Kojima, M. Latroche, H. Li, M. V. Lototskyy, J. W. Makepeace, K. T.
Møller, L. Naheed, P. Ngene, D. Noréus, M. M. Nygård, S. Orimo, M.
Conflict of Interest Paskevicius, L. Pasquini, D. B. Ravnsbæk, M. Veronica Sofianos, T. J.
Udovic, T. Vegge, G. S. Walker, C. J. Webb, C. Weidenthaler, C. Zlotea, J.
The authors declare no conflict of interest. Alloys Compd. 2020, 827, 153548.
[35] R. Ströbel, L. Jörissen, T. Schliermann, V. Trapp, W. Schütz, K.
Bohmhammel, G. Wolf, J. Garche, J. Power Sources 1999, 84, 221–224.
[36] F. L. Darkrim, P. Malbrunot, G. P. Tartaglia, Int. J. Hydrogen Energy 2002,
Keywords: energy conversion · formic acid · dehydrogenation ·
27, 193–202.
homogeneous catalysis · ligand effects [37] J. G. Vitillo, L. Regli, S. Chavan, G. Ricchiardi, G. Spoto, P. D. C. Dietzel, S.
Bordiga, A. Zecchina, J. Am. Chem. Soc. 2008, 130, 8386–8396.
[38] D. P. Broom, C. J. Webb, G. S. Fanourgakis, G. E. Froudakis, P. N.
Trikalitis, M. Hirscher, Int. J. Hydrogen Energy 2019, 44, 7768–7779.
[1] IEA World Energy Balances (all rights reserved), 2019, https://
[39] O. K. Farha, A. Özgür Yazaydın, I. Eryazici, C. D. Malliakas, B. G. Hauser,
www.iea.org/data-and-statistics/data-product/world-energy-balances
M. G. Kanatzidis, S. T. Nguyen, R. Q. Snurr, J. T. Hupp, Nat. Chem. 2010,
(accessed May 11, 2021).
2, 944–948.
[2] A. Z. Ttel, Naturwissenschaften 2004, 91, 157–172.
[40] K. M. Thomas, Catal. Today 2007, 120, 389–398.
[3] Earth’s CO2 Home Page, http://www.co2.earth (accessed Mar. 25, 2020).
[41] S. Orimo, Y. Nakamori, J. R. Eliseo, A. Züttel, C. M. Jensen, Chem. Rev.
[4] B. Bolin, B. R. Doos, Greenhouse Effect, Wiley, New York, NY (USA),
2007, 107, 4111–4132.
1989.
[42] J. Graetz, Chem. Soc. Rev. 2009, 38, 73–82.
[5] S. H. Schneider, Science 1989, 243, 771–781.
[43] P. Preuster, C. Papp, P. Wasserscheid, Acc. Chem. Res. 2017, 50, 74–85.
[6] K. Christopher, R. Dimitrios, Energy Environ. Sci. 2012, 5, 6640–6651.
[44] M. Niermann, S. Drünert, M. Kaltschmitt, K. Bonhoff, Energy Environ. Sci.
[7] P. Moriarty, D. Honnery, Renewable Sustainable Energy Rev. 2012, 16,
2019, 12, 290–307.
244–252.
[45] Q. Zhu, Q. Xu, Energy Environ. Sci. 2015, 8, 478–512.
[8] A. Azarpour, S. Suhaimi, G. Zahedi, A. Bahadori, Arabian J. Sci. Eng.
[46] M. Nielsen, E. Alberico, W. Baumann, H. Drexler, H. Junge, S. Gladiali, M.
2013, 38, 317–328.
Beller, Nature 2013, 495, 85–89.
[9] J. A. Turner, Science 1999, 285, 687–689.
[47] H. Kawanami, Y. Himeda, G. Laurenczy in Advances in Inorganic
[10] M. Momirlan, T. N. Veziroglu, Renewable Sustainable Energy Rev. 2002,
Chemistry, Vol. 70 (Eds.: R. van Eldik, C. D. Hubbard), Academic Press,
6, 141–179.
[11] S. Dunn, Int. J. Hydrogen Energy 2002, 27, 235–264. 2017, pp. 395–427.
[12] A. Midilli, M. Ay, I. Dincer, M. A. Rosen, Renewable Sustainable Energy [48] H. Jiang, S. K. Singh, J. Yan, X. Zhang, Q. Xu, ChemSusChem 2010, 3,
Rev. 2005, 9, 273–287. 541–549.
[13] A. Midilli, I. Dincer, Int. J. Hydrogen Energy 2007, 32, 511–524. [49] T. C. Johnson, D. J. Morris, M. Wills, Chem. Soc. Rev. 2010, 39, 81–88.
[14] D. Honnery, P. Moriarty, Int. J. Hydrogen Energy 2009, 34, 727–736. [50] X. Zhou, Y. Huang, W. Xing, C. Liu, J. Liao, T. Lu, Chem. Commun. 2008,
[15] B. Heid, M. Linder, A. Orthofer, M. Wilthaner, Hydrogen: The Next Wave 3540.
for Electric Vehicles?, 2017, https://www.mckinsey.com/industries/auto- [51] K. Tedsree, T. Li, S. Jones, C. W. A. Chan, K. M. K. Yu, P. A. J. Bagot, E. A.
motive-and-assembly/our-insights/hydrogen-the-next-wave-for-elec- Marquis, G. D. W. Smith, S. C. E. Tsang, Nat. Nanotechnol. 2011, 6, 302–
tric-vehicles (accessed May 11, 2021). 307.
[16] I. K. Kapdan, F. Kargi, Enzyme Microb. Technol. 2006, 38, 569–582. [52] D. L. Trimm, Appl. Catal. A 2005, 296, 1–11.
[17] L. Schlapbach, A. Zuttel, Nature 2001, 414, 353–358. [53] M. Iguchi, M. Chatterjee, N. Onishi, Y. Himeda, H. Kawanami, Sustain.
[18] A. Midilli, M. Ay, I. Dincer, M. A. Rosen, Renewable Sustainable Energy Energy Fuels 2018, 2, 1719–1725.
Rev. 2005, 9, 255–271. [54] A. Iturmendi, L. Rubio-Pérez, J. J. Pérez-Torrente, M. Iglesias, L. A. Oro,
[19] P. Leone, A. Lanzini, P. Squillari, P. Asinari, M. Santarelli, R. Borchiellini, Organometallics 2018, 37, 3611–3618.
M. Calì, Int. J. Hydrogen Energy 2008, 33, 3167–3172. [55] A. Boddien, F. Gärtner, C. Federsel, P. Sponholz, D. Mellmann, R.
[20] G. W. Crabtree, M. S. Dresselhaus, MRS Bull. 2008, 33, 421–428. Jackstell, H. Junge, M. Beller, Angew. Chem. Int. Ed. 2011, 50, 6411–
[21] P. Leone, A. Lanzini, M. Santarelli, M. Calì, F. Sagnelli, A. Boulanger, A. 6414; Angew. Chem. 2011, 123, 6535–6538.
Scaletta, P. Zitella, J. Power Sources 2010, 195, 239–248. [56] G. Papp, J. Csorba, G. Laurenczy, F. Joó, Angew. Chem. Int. Ed. 2011, 50,
[22] U. Lucia, Renewable Sustainable Energy Rev. 2014, 30, 164–169. 10433–10435; Angew. Chem. 2011, 123, 10617–10619.
[23] Y. Haseli, Int. J. Hydrogen Energy 2018, 43, 9015–9021. [57] H. Horváth, G. Papp, R. Szabolcsi, Á. Kathó, F. Joó, ChemSusChem 2015,
[24] C. H. L. G. European, C. T. Fuel, C. D. F. E. European, C. D. G. F. Euro- 8, 3036–3038.
pean, Hydrogen Energy and Fuel Cells: A Vision of Our Future:[final [58] J. Kothandaraman, M. Czaun, A. Goeppert, R. Haiges, J. Jones, R. B.
Report of the High Level Group]., Vol. 20719, European Commission, May, G. K. S. Prakash, G. A. Olah, ChemSusChem 2015, 8, 1442–1451.
Directorate General for Research, 2003. [59] K. Sordakis, A. F. Dalebrook, G. Laurenczy, ChemCatChem 2015, 7,
[25] M. Wang, U. S. Department of Energy, Well-to-wheel energy use and 2332–2339.
greenhouse gas emissions of advanced fuel/vehicle systems North [60] H. Kawanami, Y. Himeda, G. Laurenczy in Advances in Inorganic
American analysis, 2001, https://doi.org/10.2172/781268. https:// Chemistry, Vol. 70 (Eds.: R. van Eldik, C. D. Hubbard), Academic Press,
www.osti.gov/servlets/purl/781268. Cambridge, MA, 2017, pp. 395–427.
[26] K. Sordakis, C. Tang, L. K. Vogt, H. Junge, P. J. Dyson, M. Beller, G. [61] M. Iguchi, Y. Himeda, Y. Manaka, K. Matsuoka, H. Kawanami,
Laurenczy, Chem. Rev. 2018, 118, 372–433. ChemCatChem 2016, 8, 886–890.
[27] S. Nojima, Encyclopedia of Electrochemical Power Sources 2009, 39, [62] L. Wang, Q. Han, S. Hu, D. Li, P. Zhang, S. Chen, J. Xu, B. Liu, Appl. Catal.
414–420. B 2015, 164, 128–134.
[28] L. Schlapbach, A. Zuttel, Nature 2001, 414, 353–358. [63] M. Czaun, J. Kothandaraman, A. Goeppert, B. Yang, S. Greenberg, R. B.
[29] S. Gursu, S. A. Sherif, T. N. Veziroglu, J. W. Sheffield, J. Energy Resour. May, G. A. Olah, G. K. S. Prakash, ACS Catal. 2016, 6, 7475–7484.
Technol. 1993, 115, 221–227. [64] Z. Li, Q. Xu, Acc. Chem. Res. 2017, 50, 1449–1458.
[30] U. Eberle, M. Felderhoff, F. Schüth, Angew. Chem. Int. Ed. 2009, 48, [65] L. Han, L. Zhang, H. Wu, H. Zu, P. Cui, J. Guo, R. Guo, J. Ye, J. Zhu, X.
6608–6630; Angew. Chem. 2009, 121, 6732–6757. Zheng, L. Yang, Y. Zhong, S. Liang, L. Wang, Adv. Sci. 2019, 6, 1900006.
[31] A. W. C. van den Berg, C. O. Areán, Chem. Commun. 2008, 668–681. [66] C. Fellay, P. J. Dyson, G. Laurenczy, Angew. Chem. Int. Ed. 2008, 47,
[32] J. G. Vitillo, G. Ricchiardi, G. Spoto, A. Zecchina, Phys. Chem. Chem. 3966–3968; Angew. Chem. 2008, 120, 4030–4032.
Phys. 2005, 7, 3948–3954. [67] B. Loges, A. Boddien, H. Junge, M. Beller, Angew. Chem. Int. Ed. 2008,
[33] M. Mohan, V. K. Sharma, E. A. Kumar, V. Gayathri, Energy Storage 2019, 47, 3962–3965; Angew. Chem. 2008, 120, 4026–4029.
1, e35. [68] R. S. Coffey, Chem. Commun. 1967, 923.

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2679 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

[69] M. Grasemann, G. Laurenczy, Energy Environ. Sci. 2012, 5, 8171–8181. [111] D. Mellmann, P. Sponholz, H. Junge, M. Beller, Chem. Soc. Rev. 2016,
[70] E. Fujita, J. T. Muckerman, Y. Himeda, Biochim. Biophys. Acta Bioenerg. 45, 3954–3988.
2013, 1827, 1031–1038. [112] W. Wang, Y. Himeda, J. T. Muckerman, G. F. Manbeck, E. Fujita, Chem.
[71] X. Wang, Q. Meng, L. Gao, Z. Jin, J. Ge, C. Liu, W. Xing, Int. J. Hydrogen Rev. 2015, 115, 12936–12973.
Energy 2018, 43, 7055–7071. [113] C. Guan, Y. Pan, T. Zhang, M. J. Ajitha, K. W. Huang, Chem. Asian J.
[72] A. Léval, A. Agapova, C. Steinlechner, E. Alberico, H. Junge, M. Beller, 2020, 15, 937–946.
Green Chem. 2020, 22, 913–920. [114] L. Piccirilli, D. Lobo Justo Pinheiro, M. Nielsen, Catalysts 2020, 10, 773.
[73] A. Boddien, B. Loges, F. Gärtner, C. Torborg, K. Fumino, H. Junge, R. [115] M. Iglesias, L. A. Oro, Eur. J. Inorg. Chem. 2018, 2125–2138.
Ludwig, M. Beller, J. Am. Chem. Soc. 2010, 132, 8924–8934. [116] J. Eppinger, K. Huang, ACS Energy Lett. 2016, 2, 188–195.
[74] R. Langer, Y. Diskin-Posner, G. Leitus, L. J. W. Shimon, Y. Ben-David, D. [117] N. Onishi, M. Iguchi, X. Yang, R. Kanega, H. Kawanami, Q. Xu, Y.
Milstein, Angew. Chem. Int. Ed. 2011, 50, 9948–9952; Angew. Chem. Himeda, Adv. Energy Mater. 2019, 9, 1801275.
2011, 123, 10122–10126. [118] Hong Wang, Master thesis, Dalian University of Technology (China),
[75] A. Boddien, F. Gärtner, R. Jackstell, H. Junge, A. Spannenberg, W. 2017.
Baumann, R. Ludwig, M. Beller, Angew. Chem. Int. Ed. 2010, 49, 8993– [119] E. N. Yurtchenko, N. P. Anikeenko, React. Kinet. Catal. Lett. 1975, 2, 65–
8996; Angew. Chem. 2010, 122, 9177–9181. 72.
[76] A. Boddien, D. Mellmann, F. Gaertner, R. Jackstell, H. Junge, P. J. Dyson, [120] S. H. Strauss, K. H. Whitmire, D. F. Shriver, J. Organomet. Chem. 1979,
G. Laurenczy, R. Ludwig, M. Beller, Science 2011, 333, 1733–1736. 174, C59-C62.
[77] X. Yang, Dalton Trans. 2013, 42, 11987. [121] B. The Khai, A. Arcelli, J. Organomet. Chem. 1986, 309, C63–C66.
[78] T. Zell, B. Butschke, Y. Ben-David, D. Milstein, Chem. Eur. J. 2013, 19, [122] Y. Gao, J. K. Kuncheria, H. A. Jenkins, R. J. Puddephatt, G. P. A. Yap, J.
8068–8072. Chem. Soc. Dalton Trans. 2000, 3212–3217.
[79] E. A. Bielinski, P. O. Lagaditis, Y. Zhang, B. Q. Mercado, C. Würtele, W. H. [123] C. Fellay, N. Yan, P. J. Dyson, G. Laurenczy, Chem. Eur. J. 2009, 15,
Bernskoetter, N. Hazari, S. Schneider, J. Am. Chem. Soc. 2014, 136, 3752–3760.
10234–10237. [124] W. Gan, C. Fellay, P. J. Dyson, G. Laurenczy, J. Coord. Chem. 2010, 63,
[80] F. Bertini, I. Mellone, A. Ienco, M. Peruzzini, L. Gonsalvi, ACS Catal. 2685–2694.
2015, 5, 1254–1265. [125] W. Gan, D. J. M. Snelders, P. J. Dyson, G. Laurenczy, ChemCatChem
[81] L. Wang, H. Sun, Z. Zuo, X. Li, W. Xu, R. Langer, O. Fuhr, D. Fenske, Eur. 2013, 5, 1126–1132.
J. Inorg. Chem. 2016, 5205–5214. [126] A. Guerriero, H. Bricout, K. Sordakis, M. Peruzzini, E. Monflier, F. Hapiot,
[82] M. Montandon-Clerc, A. F. Dalebrook, G. Laurenczy, J. Catal. 2016, 343, G. Laurenczy, L. Gonsalvi, ACS Catal. 2014, 4, 3002–3012.
62–67. [127] S. Moret, P. J. Dyson, G. Laurenczy, Dalton Trans. 2013, 42, 4353–4356.
[83] J. B. Curley, N. E. Smith, W. H. Bernskoetter, N. Hazari, B. Q. Mercado, [128] W. Gan, P. J. Dyson, G. Laurenczy, React. Kinet. Catal. Lett. 2009, 98,
Organometallics 2018, 37, 3846–3853. 205–213.
[129] V. Henricks, I. Yuranov, N. Autissier, G. Laurenczy, Catalysts 2017, 7,
[84] W. Zhou, Z. Wei, A. Spannenberg, H. Jiao, K. Junge, H. Junge, M. Beller,
348.
Chem. Eur. J. 2019, 25, 8459–8464.
[130] A. Boddien, B. Loges, H. Junge, M. Beller, ChemSusChem 2008, 1, 751–
[85] J. H. Shin, D. G. Churchill, G. Parkin, J. Organomet. Chem. 2002, 642, 9–
758.
15.
[131] A. Boddien, B. Loges, H. Junge, F. Gärtner, J. R. Noyes, M. Beller, Adv.
[86] M. C. Neary, G. Parkin, Chem. Sci. 2015, 6, 1859–1865.
Synth. Catal. 2009, 351, 2517–2520.
[87] G. N. Khairallah, R. A. J. O. Hair, Int. J. Mass Spectrom. 2006, 254, 145–
[132] H. Junge, A. Boddien, F. Capitta, Tetrahedron Lett. 2009, 50, 1603–1606.
151.
[133] B. Loges, A. Boddien, H. Junge, J. R. Noyes, W. Baumann, M. Beller,
[88] T. W. Myers, L. A. Berben, Chem. Sci. 2014, 5, 2771–2777.
Chem. Commun. 2009, 4185–4187.
[89] Q. Q. Lu, H. Z. Yu, Y. Fu, Chem. Eur. J. 2016, 22, 4584–4591.
[134] C. Prichatz, M. Trincado, L. Tan, F. Casas, A. Kammer, H. Junge, M.
[90] C. Chauvier, A. Tlili, C. Das Neves Gomes, P. Thuéry, T. Cantat, Chem.
Beller, H. Grützmacher, ChemSusChem 2018, 11, 3092–3095.
Sci. 2015, 6, 2938–2942.
[135] M. Czaun, A. Goeppert, J. Kothandaraman, R. B. May, R. Haiges, G. K. S.
[91] M. Vogt, A. Nerush, Y. Diskin-Posner, Y. Ben-David, D. Milstein, Chem.
Prakash, G. A. Olah, ACS Catal. 2014, 4, 311–320.
Sci. 2014, 5, 2043–2051. [136] G. Papp, G. Olveti, H. Horvath, A. Katho, F. Joo, Dalton Trans. 2016, 45,
[92] M. A. Esteruelas, C. García-Yebra, J. Martín, E. Oñate, ACS Catal. 2018, 8, 14516–14519.
11314–11323. [137] S. Oldenhof, B. de Bruin, M. Lutz, M. A. M. Siegler, F. W. Patureau, J. I.
[93] J. Shin, J. H. Yoon, S. H. Lee, T. H. Park, Bioresour. Technol. 2010, 101, van der Vlugt, J. N. H. Reek, Chem. Eur. J. 2013, 19, 11507–11511.
S53–S58. [138] S. Oldenhof, M. Lutz, B. de Bruin, J. I. van der Vlugt, J. N. H. Reek, Chem.
[94] Y. Himeda, Green Chem. 2009, 11, 2018–2022. Sci. 2015, 6, 1027–1034.
[95] Y. Maenaka, T. Suenobu, S. Fukuzumi, Energy Environ. Sci. 2012, 5, [139] S. Fukuzumi, T. Kobayashi, T. Suenobu, ChemSusChem 2008, 1, 827–
7360. 834.
[96] J. H. Barnard, C. Wang, N. G. Berry, J. Xiao, Chem. Sci. 2013, 4, 1234. [140] Y. Himeda, N. Onozawa-Komatsuzaki, H. Sugihara, H. Arakawa, K.
[97] Z. Wang, S. Lu, J. Li, J. Wang, C. Li, Chem. Eur. J. 2015, 21, 12592–12595. Kasuga, Organometallics 2004, 23, 1480–1483.
[98] Y. Gao, J. Kuncheria, R. J. Puddephatt, G. P. A. Yap, Chem. Commun. [141] Y. Himeda, N. Onozawa-Komatsuzaki, H. Sugihara, K. Kasuga, J. Am.
1998, 2365–2366. Chem. Soc. 2005, 127, 13118–13119.
[99] Z. Wang, S. Lu, J. Wu, C. Li, J. Xiao, Eur. J. Inorg. Chem. 2016, 490–496. [142] Y. Himeda, N. Onozawa-Komatsuzaki, H. Sugihara, K. Kasuga, Organo-
[100] X. Li, X. Ma, F. Shi, Y. Deng, ChemSusChem 2010, 3, 71–74. metallics 2007, 26, 702–712.
[101] J. D. Scholten, M. H. G. Prechtl, J. Dupont, ChemCatChem 2010, 2, [143] Y. Himeda, N. Onozawa-Komatsuzaki, S. Miyazawa, H. Sugihara, T.
1265–1270. Hirose, K. Kasuga, Chem. Eur. J. 2008, 14, 11076–11081.
[102] Y. Yasaka, C. Wakai, N. Matubayasi, M. Nakahara, J. Phys. Chem. A 2010, [144] Y. Himeda, Eur. J. Inorg. Chem. 2007, 3927–3941.
114, 3510–3515. [145] Y. Himeda, S. Miyazawa, T. Hirose, ChemSusChem 2011, 4, 487–493.
[103] M. E. M. Berger, D. Assenbaum, N. Taccardi, E. Spiecker, P. Wassersc- [146] J. F. Hull, Y. Himeda, W. Wang, B. Hashiguchi, R. Periana, D. J. Szalda,
heid, Green Chem. 2011, 13, 1411. J. T. Muckerman, E. Fujita, Nat. Chem. 2012, 4, 383–388.
[104] S. Enthaler, J. von Langermann, T. Schmidt, Energy Environ. Sci. 2010, 3, [147] T. Wonglakhon, P. Surawatanawong, Dalton Trans. 2018, 47, 17020–
1207–1217. 17031.
[105] W. H. Bernskoetter, N. Hazari, Acc. Chem. Res. 2017, 50, 1049–1058. [148] N. Johnee Britto, A. S. Rajpurohit, K. Jagan, M. Jaccob, J. Phys. Chem. C
[106] N. Onishi, G. Laurenczy, M. Beller, Y. Himeda, Coord. Chem. Rev. 2018, 2019, 123, 25061–25073.
373, 317–332. [149] W. Wang, J. F. Hull, J. T. Muckerman, E. Fujita, T. Hirose, Y. Himeda,
[107] A. K. Singh, S. Singh, A. Kumar, Catal. Sci. Technol. 2015, 6, 12–14. Chem. Eur. J. 2012, 18, 9397–9404.
[108] B. Loges, A. Boddien, F. Gärtner, H. Junge, M. Beller, Top. Catal. 2010, [150] W. H. Wang, S. Xu, Y. Manaka, Y. Suna, H. Kambayashi, J. T. Muckerman,
53, 902–914. E. Fujita, Y. Himeda, ChemSusChem 2014, 7, 1976–1983.
[109] M. Czaun, A. Goeppert, R. May, R. Haiges, G. K. S. Prakash, G. A. Olah, [151] P. Segura, J. Org. Chem. 1985, 50, 1045–1053.
ChemSusChem 2011, 4, 1241–1248. [152] S. Siek, D. B. Burks, D. L. Gerlach, G. Liang, J. M. Tesh, C. R. Thompson,
[110] C. Fink, M. Montandon-Clerc, G. Laurenczy, Chimia 2015, 69, 746–752. F. Qu, J. E. Shankwitz, R. M. Vasquez, N. Chambers, G. J. Szulczewski,

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2680 © 2021 Wiley-VCH GmbH
1864564x, 2021, 13, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cssc.202100602 by Dalian Institute Of Chemical, Wiley Online Library on [07/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
ChemSusChem doi.org/10.1002/cssc.202100602

D. B. Grotjahn, C. E. Webster, E. T. Papish, Organometallics 2017, 36, [182] P. Sponholz, D. Mellmann, H. Junge, M. Beller, ChemSusChem 2013, 6,
1091–1106. 1172–1176.
[153] Y. Manaka, W. Wang, Y. Suna, H. Kambayashi, J. T. Muckerman, E. [183] I. Mellone, F. Bertini, M. Peruzzini, L. Gonsalvi, Catal. Sci. Technol. 2016,
Fujita, Y. Himeda, Catal. Sci. Technol. 2014, 4, 34–37. 6, 6504–6512.
[154] W. Wang, M. Z. Ertem, S. Xu, N. Onishi, Y. Manaka, Y. Suna, H. [184] Z. Xin, J. Zhang, K. Sordakis, M. Beller, C. Du, G. Laurenczy, Y. Li,
Kambayashi, J. T. Muckerman, E. Fujita, Y. Himeda, ACS Catal. 2015, 5, ChemSusChem 2018, 11, 2077–2082.
5496–5504. [185] O. Fuganti, J. P. Da Silva, D. F. Back, M. P. de Araujo, J. Mol. Struct. 2020,
[155] N. Onishi, M. Z. Ertem, S. Xu, A. Tsurusaki, Y. Manaka, J. T. Muckerman, 1200, 127129.
E. Fujita, Y. Himeda, Catal. Sci. Technol. 2016, 6, 988–992. [186] S. Hsu, S. Rommel, P. Eversfield, K. Muller, E. Klemm, W. R. Thiel, B.
[156] W. Wang, H. Wang, Y. Yang, X. Lai, Y. Li, J. Wang, Y. Himeda, M. Bao, Plietker, Angew. Chem. Int. Ed. 2014, 53, 7074–7078; Angew. Chem.
ChemSusChem 2020, 13, 5015–5022. 2014, 126, 7194–7198.
[157] M. Iguchi, Y. Himeda, Y. Manaka, H. Kawanami, ChemSusChem 2016, 9, [187] S. Hsu, B. Plietker, ChemCatChem 2013, 5, 126–129.
2749–2753. [188] L. S. Jongbloed, B. De Bruin, J. N. H. Reek, M. Lutz, J. I. Van Der Vlugt,
[158] M. Iguchi, H. Zhong, Y. Himeda, H. Kawanami, Chem. Eur. J. 2017, 23, Catal. Sci. Technol. 2016, 6, 1320–1327.
17017–17021. [189] Y. Pan, C. Pan, Y. Zhang, H. Li, S. Min, X. Guo, B. Zheng, H. Chen, A.
[159] R. Kanega, N. Onishi, L. Wang, K. Murata, J. T. Muckerman, E. Fujita, Y. Anders, Z. Lai, J. Zheng, K. Huang, Chem. Asian J. 2016, 11, 1357–1360.
Himeda, Chem. Eur. J. 2018, 24, 18389–18392. [190] S. Cohen, V. Borin, I. Schapiro, S. Musa, S. De-Botton, N. V. Belkova, D.
[160] N. Onishi, R. Kanega, E. Fujita, Y. Himeda, Adv. Synth. Catal. 2019, 361, Gelman, ACS Catal. 2017, 7, 8139–8146.
289–296. [191] L. Rubio-Perez, M. Iglesias, J. Munarriz, V. Polo, V. Passarelli, J. J. Perez-
[161] M. Iguchi, N. Onishi, Y. Himeda, H. Kawanami, ChemPhysChem 2019, Torrente, L. A. Oro, Chem. Sci. 2017, 8, 4811–4822.
20, 1296–1300. [192] A. Iturmendi, M. Iglesias, J. Munarriz, V. Polo, V. Passarelli, J. J. Perez-
[162] H. Kawanami, M. Iguchi, Y. Himeda, Inorg. Chem. 2020, 59, 4191–4199. Torrente, L. A. Oro, Green Chem. 2018, 20, 4875–4879.
[163] J. Li, J. Li, D. Zhang, C. Liu, ACS Catal. 2016, 6, 4746–4754. [193] J. Kothandaraman, M. Czaun, A. Goeppert, R. Haiges, J. P. Jones, R. B.
[164] S. Lu, Z. Wang, J. Wang, J. Li, C. Li, Green Chem. 2018, 20, 1835–1840. May, G. K. S. Prakash, G. A. Olah, ChemSusChem 2015, 8, 1442–1451.
[165] A. Matsunami, Y. Kayaki, T. Ikariya, Chem. Eur. J. 2015, 21, 13513– [194] E. Alberico, A. J. J. Lennox, L. K. Vogt, H. Jiao, W. Baumann, H. Drexler,
13517. M. Nielsen, A. Spannenberg, M. P. Checinski, H. Junge, M. Beller, J. Am.
[166] A. Matsunami, S. Kuwata, Y. Kayaki, ACS Catal. 2017, 7, 4479–4484. Chem. Soc. 2016, 138, 14890–14904.
[167] C. Fink, G. Laurenczy, Dalton Trans. 2017, 46, 1670–1676. [195] A. Agapova, E. Alberico, A. Kammer, H. Junge, M. Beller, ChemCatChem
[168] C. Fink, L. Chen, G. Laurenczy, Z. Anorg. Allg. Chem. 2018, 644, 740– 2019, 11, 1910–1914.
744. [196] W. Baratta, E. Herdtweck, K. Siega, M. Toniutti, P. Rigo, Organometallics
[169] C. Fink, G. Laurenczy, Eur. J. Inorg. Chem. 2019, 2381–2387. 2005, 24, 1660–1669.
[170] G. Menendez Rodriguez, C. Domestici, A. Bucci, M. Valentini, C. [197] A. Léval, H. Junge, M. Beller, Eur. J. Inorg. Chem. 2020, 1293–1299.
Zuccaccia, A. Macchioni, Eur. J. Inorg. Chem. 2018, 2247–2250. [198] S. Kar, M. Rauch, G. Leitus, Y. Ben-David, D. Milstein, Nat. Catal. 2021, 4,
[171] S. C. Fischmeister, H. Huang, V. Dorcet, T. Roisnel, C. Bruneau, C. 193–201.
Fischmeister, Organometallics 2017, 36, 3152–3162. [199] J. Hannedouche, G. J. Clarkson, M. Wills, J. Am. Chem. Soc. 2004, 126,
[172] S. Wang, H. Huang, C. Bruneau, C. Fischmeister, ChemSusChem 2017, 986–987.
10, 4150–4154. [200] A. M. Hayes, D. J. Morris, G. J. Clarkson, M. Wills, J. Am. Chem. Soc.
[173] S. Wang, H. Huang, T. Roisnel, C. Bruneau, C. Fischmeister, ChemSu- 2005, 127, 7318–7319.
sChem 2019, 12, 179–184. [201] D. S. Matharu, D. J. Morris, A. M. Kawamoto, G. J. Clarkson, M. Wills,
[174] D. Ying, S. Yang-Bin, Z. Yu-Lu, N. Fan-Di, Y. Liu-Ming, Z. Xiao-Chun, Org. Lett. 2005, 7, 5489–5491.
Chin. Chem. Lett. 2017, 1746–1750. [202] D. J. Morris, G. J. Clarkson, M. Wills, Organometallics 2009, 28, 4133–
[175] J. J. A. Celaje, Z. Lu, E. A. Kedzie, N. J. Terrile, J. N. Lo, T. J. Williams, Nat. 4140.
Commun. 2016, 7, 11308. [203] A. Majewski, D. J. Morris, K. Kendall, M. Wills, ChemSusChem 2010, 3,
[176] P. J. Lauridsen, Z. Lu, J. J. A. Celaje, E. A. Kedzie, T. J. Williams, Dalton 431–434.
Trans. 2018, 47, 13559–13564. [204] M. L. Man, Z. Zhou, S. M. Ng, C. P. Lau, Dalton Trans. 2003, 3727–3735.
[177] C. Guan, D. Zhang, Y. Pan, M. Iguchi, M. J. Ajitha, J. Hu, H. Li, C. Yao, M. [205] S. Fukuzumi, T. Kobayashi, T. Suenobu, J. Am. Chem. Soc. 2010, 132,
Huang, S. Ming, J. Zheng, Y. Himeda, H. Kawanami, K. Huang, Inorg. 1496–1497.
Chem. 2017, 56, 438–445. [206] S. J. S. M. Samec, J. Bäckvall, P. G. Andersson, P. Brandt, Chemical
[178] S. Patra, M. K. Awasthi, R. K. Rai, H. Deka, S. M. Mobin, S. K. Singh, Eur. J. Society Reviews 2006, 35, 237.
Inorg. Chem. 2019, 1046–1053.
[179] R. Tanaka, M. Yamashita, L. W. Chung, K. Morokuma, K. Nozaki,
Organometallics 2011, 30, 6742–6750.
[180] I. Mellone, M. Peruzzini, L. Rosi, D. Mellmann, H. Junge, M. Beller, L.
Gonsalvi, Dalton Trans. 2013, 42, 2495–2501. Manuscript received: March 23, 2021
[181] G. Manca, I. Mellone, F. Bertini, M. Peruzzini, L. Rosi, D. Mellmann, H. Revised manuscript received: April 29, 2021
Junge, M. Beller, A. Ienco, L. Gonsalvi, Organometallics 2013, 32, 7053– Accepted manuscript online: May 7, 2021
7064. Version of record online: June 7, 2021

ChemSusChem 2021, 14, 2655 – 2681 www.chemsuschem.org 2681 © 2021 Wiley-VCH GmbH

You might also like