You are on page 1of 34

CHAPTER SIX

Interconversion of CO2/H2 and


Formic Acid Under Mild
Conditions in Water: Ligand
Design for Effective Catalysis
Wan-Hui Wang*,†, Yuichiro Himeda*,†, James T. Muckerman{,
and Etsuko Fujita{
*National Institute of Advanced Industrial Science and Technology, Tsukuba, Ibaraki, Japan

Japan Science and Technology Agency, Kawaguchi, Saitama, Japan
{
Chemistry Department, Brookhaven National Laboratory, Upton, New York, USA

Contents
1. Introduction 190
2. Hydrogenation of CO2 to Formic Acid 193
2.1 Historical background 193
2.2 Design and synthesis of complexes with proton-responsive ligands 198
2.3 Mechanism of catalyst activation 201
2.4 pH-dependent water solubility and catalyst recycling 209
3. Dehydrogenation of Formic Acid 211
3.1 Historical background 211
3.2 pH-dependent activity 213
3.3 Electronic effect for catalyst activation 214
3.4 Generation of high-pressure H2 for practical use 215
4. Reversible Hydrogen Storage by Interconversion of CO2/H2 and HCO2H 217
5. Concluding Remarks 219
Acknowledgments 220
References 220

Abstract
Recent significant progress in the homogeneous catalytic hydrogenation of CO2 to for-
mate (the conjugate base of formic acid) and dehydrogenation of formic acid in various
solvents including water is summarized. While formic acid is not the perfect H2 storage
solution, many researchers consider it better than other methods at this time because
the interconversion of CO2 and formic acid can take place cleanly to form H2 without
detectable CO under mild conditions. In this chapter, we explain how inspirations from
biological systems guide us to design homogeneous transition-metal catalysts for

Advances in Inorganic Chemistry, Volume 66 # 2014 Elsevier Inc. 189


ISSN 0898-8838 All rights reserved.
http://dx.doi.org/10.1016/B978-0-12-420221-4.00006-8
190 Wan-Hui Wang et al.

carrying out the interconversion of CO2 and formate under ambient conditions in envi-
ronmentally benign and economically desirable water solvent.
Keywords: CO2 hydrogenation, Dehydrogenation of formic acid, Ir complexes,
Proton-responsive ligands, H2 storage, Bioinspired catalysts

1. INTRODUCTION
The capture and utilization of CO2 as fuels and chemicals are impor-
tant scientific projects in view of growing fear of depletion of fossil fuels and
global warming. The CO2 concentration levels in the atmosphere are dra-
matically rising owing primarily to the human activities of burning fossil
fuels and deforestation during the past 50 years, and have now reached
almost 400 ppm. Research in the fields of CO2 mitigation and utilization
has received significant attention; however, because CO2 is such a stable
molecule, there remain great chemical, physical, and engineering challenges
to convert CO2 into industrially important chemicals and fuels. Because
CO2 is the end product of hydrocarbon combustion (H2O is the coproduct),
energy is needed to convert it to fuels such as CO and formic acid (FA). CO2
can be reduced by photochemical methods (i.e., artificial photosynthesis),
electrochemical reduction using photovoltaic electricity, and thermal
hydrogenation using photoproduced hydrogen. While one-electron reduc-
tion of CO2 to CO2  is thermodynamically unfavorable (Equation 6.1),
multielectron proton-coupled reactions can take place at more positive
potentials (Equations 6.2 and 6.3):
CO2 þ e ! CO2 E 0 ¼ 1:90 V ð6:1Þ
þ  0
CO2 þ 2H þ 2e ! CO þ H2 O E ¼ 0:53 V ð6:2Þ
þ  0
CO2 þ 2H þ 2e ! HCO2 H E ¼ 0:61 V ð6:3Þ
Here, E0 (as opposed to Eo) is the formal potential versus the normal
hydrogen electrode (NHE) at standard conditions in aqueous solution at
25  C, 1 atm of gases, and 1 M solutes, but pH 7 instead of pH 0.
During the last 25 years, we have investigated the kinetics and mechanisms
of photochemical CO2 reduction using (1) metal macrocycles; (2)
Ru(bpy)2(CO)Xnþ (bpy ¼ 2,20 -bipyridine, X ¼ Cl, n ¼ 1; X ¼ CO, n ¼ 2);
and (3) Re(dmb)(CO)3Cl or [Re(dmb)(CO)3]2 (dmb ¼ 4,40 -dimethyl-2,20 -
bipyridine) under various conditions (1–8). However, the turnover frequency
(TOF) and the turnover number (TON) for CO or formate production are
Ligand Design for Effective Catalysis 191

rather limited. This raises the question of whether direct photochemical CO2
reduction is better than thermal CO2 hydrogenation. Photo- and electro-
chemical CO2 reduction systems, and biological systems, directly utilize pro-
tons and electrons for fuel (e.g., formate) synthesis from CO2. While CO2
hydrogenation in water is of interest, it is rather complicated owing to the
acid–base equilibrium of CO2 as shown in Equation (6.4). Although the
hydrogenation of CO2 into FA (Equation 6.5) in the gas phase is endergonic
(DG 298 ¼ þ33 kJ mol1), the reaction in the aqueous phase is exergonic
(DG 298 ¼ –4 kJ mol1), and the presence of a base makes the hydrogenation
of CO2 more favorable. Similarly, on the basis of theoretical calculations, the
hydrogenation of bicarbonate into formate in water (Equation 6.6) is believed
to be exergonic. Formate is the conjugate base of FA (Equation 6.7) and the
dehydrogenation is more favorable in acidic conditions.
pK1 = 6.35 pK2 = 10.33
CO 2 + H2O H2CO3 HCO3− + H+ CO32− + 2H+ ð6:4Þ

CO2 + H2 HCO2H ð6:5Þ

HCO3– + H2 HCO2– + H2O ð6:6Þ

pKa = 3.75
HCO2H HCO2− + H+ ð6:7Þ

In this chapter, we will summarize recent investigations of homogeneous


catalytic hydrogenation of CO2 to formate and dehydrogenation of FA in
various solvents including water. The term “hydrogenation of CO2” is fre-
quently used in this book chapter and elsewhere, but such reactions in basic
aqueous solutions may involve HCO3  or CO3 2 as substrates depending
on the pH of the solution. In some cases, “hydrogenation of CO2” can be
done in a HCO3  or CO3 2 solution in the absence of additional CO2, but
it is known that such reactions do not always work.
Formate/FA can be used not only for fuel cells but also as an H2 storage
medium. While FA is not the perfect H2 storage solution (its principal draw-
back being that it stores only 4.4 wt% of H2), many researchers consider it
better than other methods at this time since the conversion of CO2 and FA
can take place cleanly to form H2 without detectable CO under mild con-
ditions. In addition, FA is used as a preservative, insecticide, and industrial
material for synthetic processes.
Recently, Crabtree (9) and others (10–12) published excellent reviews on
ligand design with additional functional groups such as proton-responsive
192 Wan-Hui Wang et al.

ligands capable of gaining or losing one or more protons, ligands having a


hydrogen-bonding function, electroresponsive ligands capable of gaining or
losing one or more electrons, photoresponsive ligands capable of undergoing
a useful change in properties upon irradiation, and hemilabile ligands capable
of providing a vacant coordination site. Biological systems cleverly use
proton-responsive ligands, hydrogen-bonding interactions, and pendent
bases in the second coordination sphere. For example, based on the structure
of Fe–Fe hydrogenases, H2 activation (i.e., the Fe center accepting a hydride
and a pendent nitrogen base accepting a proton) has been proposed to occur
via addition of H2 to a metal center followed by heterolysis of the bound H2
(Scheme 6.1). Various bioinspired model complexes have been developed,
and an advantageous effect of a pendent base has been proved for H2 oxidation
and H2 production (13).
Formate dehydrogenases containing tungsten or molybdenum are
enzymes that catalyze the oxidation of formate to CO2. The most com-
mon class of the enzymes directly transfers a hydride moiety from formate
to a cation of nicotinamide adenine dinucleotide phosphate (NAHPHþ);
however, the reverse reaction is difficult to drive because the reduction
potential of NADPþ is more positive than that of CO2. The formate
dehydrogenases that contain molybdenum or tungsten cofactors can trans-
fer an electron from formate to reduce quinone, protons, or NADPþ;
therefore, formate becomes CO2 and Hþ. In fact, interconversion of
CO2 and formate is essential to the metabolism of several bacteria. Reda
et al. demonstrated reversible interconversion of CO2 and formate using
the tungsten-containing formate dehydrogenase enzyme immobilized on
an electrode surface (14). They found that it catalyzes efficient electro-
chemical reduction of CO2 to formate with a rate more than two orders
of magnitude faster than any known catalysts for the same reaction. Elec-
trochemical formate oxidation (3380 s1 at pH 8) is much faster than CO2
reduction (280 s1 at pH 7.5). Unfortunately, the structure of the enzyme
they used is not known; however, structures of several molybdenum- or

N N N H
H
S S S S S S
H2 H H
S S S CO
Fe Fe CO Fe Fe CO Fe Fe
NC NC NC
CN CN CN
OC C OC C OC C
O O O

Scheme 6.1 Proposed mechanism of H2 oxidation and production by Fe–Fe


hydrogenases.
Ligand Design for Effective Catalysis 193

tungsten-containing enzymes have been determined. In the active site,


molybdenum or tungsten is coordinated to the cis-dithiolene group of
one or two pyranopterins in addition to terminal oxo/hydroxo groups
and/or sulfido groups or side chains of serine, cysteine, selenocysteine,
or aspartate residues in a diversity of arrangements. For CO2–formate con-
version, the arginine residue is proposed to orient CO2 and formate suit-
ably for proton delivery and removal, respectively, via hydrogen-bonding
interactions with a histidine residue in the active site.
Here, we explain how biological inspirations help us to design homoge-
neous transition-metal catalysts for carrying out the interconversion of CO2
and formate under ambient conditions in environmentally benign and eco-
nomically desirable water solvent. By reviewing the results of interconver-
sion of CO2 and FA using our catalysts and catalysts from other published
studies, we hope to demonstrate a new design principle that greatly improves
the catalytic activity.

2. HYDROGENATION OF CO2 TO FORMIC ACID


2.1. Historical background
The pioneering work by Inoue et al. (15) using phosphine complexes of Ru,
Rh, Ir, etc., has opened a new avenue of homogeneous catalytic hydrogenation
of CO2 to FA. Following that work, complexes based on transition metals such
as Pd, Ni, Rh, Ru, Ir, and various ligands have been developed and utilized in
the catalysis of hydrogenation. Platinum-group metals combined with phos-
phine ligands have been demonstrated to be efficient catalysts. On the other
hand, the solvent also plays an important role in improving the catalytic effi-
ciency. Initially, phosphine complexes were widely used, but, because of their
insolubility in water, homogeneous hydrogenation of CO2 generally
proceeded in organic solvents such as DMSO despite water showing an accel-
erating effect. Noyori and Jessop et al. have achieved outstanding catalytic per-
formance (TOF up to 95,000 h1) in supercritical CO2 using RuCl(OAc)
(PMe3)4 in the presence of an amine (16–18). Homogeneous hydrogenation
of CO2 to FA in water has recently attracted increasing attention because water
is abundant, inexpensive, and eco-friendly. More importantly, hydrogenation
of CO2 in water is considerably favored (DG ¼ 4 kJ mol1) compared to the
reaction in gas phase (DG ¼ þ33 kJ mol1). Significant progress in the hydro-
genation of CO2 in aqueous media has been achieved in the last decade.
ts
en-
m-
e-
ev-
hi-
ac-
ese
th-
zed
ari-
m-
m-
su-
ve
ha-
ws
e-
vi-
re-
of
ber
m-
nu-
A-
Table 6.1 Hydrogenation of CO2 to formic acid/formate in aqueous media
Catalyst Additive P(H2/CO2) (MPa) T ( C) t (h) TON TOFa (h1) References
RhCl(TPPTS)3 NHMe2 2/2 81 0.5 – 7300 (19)
2/2 rt 12 3400 – (19)
[RuCl2(TPPMS)2]2 NaHCO3 6/3.5 80 0.03 320 9600 (30)
0.2/0.8 50 1 – 50 (30)
RuCl2(PTA)4 NaHCO3 5/0 80 – – (807) (33)
[RuCl2(C6H6)]/DPPM NaHCO3 5/3.5 70 2 2518 1259 (37)
1 b
IrH3(PNP ) KOH 4/4 200 2 300,000 150,000 (38, 39)
4/4 120 48 3,500,000 73,000
IrH3(PNP2) KOH 2.8/2.8c 185 24 348,000 14,500 (44)
3 b
FeH2(CO)(PNP ) NaOH 0.67/0.33 80 5 788 156 (45)
[Cp*Ir(4DHBP)Cl] þ
KOH 3/3 120 57 190,000 (42,000) (56)
[Cp*Ir(4DHBP)(H2O)]2þ NaHCO3 0.05/0.05 25 24 92 (7) (57)
[Cp*Ir(DHPT)Cl] þ
KOH 3/3 120 48 222,000 (33,000) (56)
[Cp*Ir(DHPT)Cl]þ KOH 0.05/0.05 30 30 81 (3.5) (56)
[Cp*Ir(6DHBP)(H2O)] 2þ
NaHCO3 0.05/0.05 25 33 330 (27) (57)
[Cp*Ir(6DHBP)(H2O)] 2þ
KHCO3 0.5/0.5 120 8 12,500 (25,200) (57)
[(Cp*Ir)2(THBPM)Cl2]2þ NaHCO3 0.05/0.05 25 336 7200 (64) (58)
[Cp*Ir(TH4BPM)(H2O)] 2þ
KHCO3 0.05/0.05 25 24 193 (66) (59)
IrI2(AcO)(bis-NHC) KOH 3/3 200 75 190,000 2500 (54)
a
The data in parenthesis are initial TOFs.
b
Solvent is H2O/THF.
c
Pressure at rt.
TPPTS, tris(3-sulfonatophenyl)phosphine); TPPMS, 3-sulfonatophenyldiphenylphosphine; PTA, 1,3,5-triaza-7-phosphaadamantane; DPPM, 1,2-bis(diphenylphosphino)
methane; PNP1, PNP2, and PNP3 (see Figure 6.1); 4DHBP, 4,40 -dihydroxyl-2,20 -bipyridine; DHPT, 4,7-dihydroxyl-1,10-phenathroline; 6DHBP, 6,60 -dihydroxyl-2,20 -
bipyridine; THBPM, 4,40 ,6,60 -tetrahydroxyl-2,20 -bipyrimidine; TH4BPM, 2,20 ,6,60 -tetrahydroxyl-4,40 -bipyrimidine; bis-NHC, N-heterocyclic carbenes (see Figure.6.2).
196 Wan-Hui Wang et al.

(19–25). The performances of most of the highly efficient complexes are listed
in Table 6.1.
In 1993, Leitner et al. reported the first water-soluble rhodium catalyst,
RhCl(TPPTS)3 (TPPTS: tris(3-sulfonatophenyl)phosphine), which gave a
high TON of 3440 under relatively mild conditions (rt, 4 MPa H2/CO2)
(26). Joó et al. performed extensive studies using phosphorous rhodium
and ruthenium complexes in amine-free aqueous solutions (27–32).
A high TOF of 9600 h1 was obtained by using [RuCl2(TPPMS)2]2
(TPPMS: 3-sulfonatophenyldiphenylphosphine) at 9.5 MPa and 80  C.
Subsequently, Laurenczy and coworkers investigated reaction mechanisms
in detail with the rhodium and ruthenium catalysts having water-soluble
1,3,5-triaza-7-phosphaadamantane (PTA) ligand (33–36). Most recently,
Beller and Laurenczy et al. reported moderate catalytic activity (TOF:
1259 h1) using in situ complex [RuCl2(C6H6)]2/DPPM (DPPM: 1,2-bis
(diphenylphosphino)methane) in aqueous NaHCO3 under 8.5 MPa of
H2/CO2 (5/3.5) at 70  C (37).
In 2009, Nozaki and coworkers developed an Ir trihydride complex
IrH3(PNP1) (1, Figure 6.1) with a PNP pincer ligand and achieved the
highest activity to that date for CO2 hydrogenation. Due to the low water
solubility of the PNP complex, THF is required as a cosolvent for the homo-
geneous catalysis. Complex 1 showed a TOF of 150,000 h1 at 200  C and a

H H H O H CO
H P H
P Ir P H O P Fe P
H N Ir H P H
N N
P N Ir H
H P
H
1 P: iPr2P 2 P: iPr2P 3 P: iPr2P 4 P: tBu2P

O O H
H H O
O O H
P N P Ir P P Ir P

N Ru CO N Ru CO X X X O X

P P
H H

5 P: tBu2P 6 P: tBu2P 7 P: tBu2P 8 P: tBu2P


N: NEt2 X: O or CH2 X: O or CH2

Figure 6.1 Pincer complexes for CO2 hydrogenation in water.


Ligand Design for Effective Catalysis 197

TON of 3,500,000 at 120  C for 48 h under 8 MPa H2/CO2 (1/1) in


H2O/THF (5/1) (38, 39). This outstanding performance attracted wide
interest soon after it was reported. The reaction mechanism of the PNP
Ir complex (39–42) and some related Co, Fe, Pd, and Ni pincer complexes
(41, 43) was investigated in computational studies. In 2011, Hazari and
coworkers investigated a IrH3(PNP2) complex (2) bearing an N–H group,
which forms stable complex 3 after reaction with CO2 (44). Their calcula-
tions indicated that CO2 insertion is facilitated by an N–H–O hydrogen
bond through an outer sphere interaction. Complex 3 achieved a maximum
TON of 348,000 and a high TOF of 18,780 h1. Milstein and coworkers
synthesized the most active iron complex, trans-[FeH2(CO)(PNP3)] (4),
which gave a TON up to 780 and TOF up to 160 h1 under low pressure
(0.6–1 MPa) in H2O/THF (10/1) (45). Almost simultaneously, Milstein
(46) and Sanford (47) reported the crystal structures of Ru PNP complex
5 and Ru PNN complex 6, which were synthesized by the reaction
of the corresponding Ru pincer complexes with CO2. These studies
suggested that the noninnocent pincer ligands play an important role in
the activation of CO2. The noninnocent character of pincer ligands is
believed to be responsible for the extraordinary performance of pincer com-
plexes in the activation of small molecules, such as H2 and CO2, through
metal–ligand cooperation (48–50). In contrast, the IrH2(PCP) pincer com-
plexes 7 were found to form k2-formato complexes 8 (51). In addition, 7 can
be utilized in the electrocatalytic reduction of CO2 to formate in
H2O/CH3CN.
Another series of water-soluble complexes 9–13 (Figure 6.2) was devel-
oped by Peris et al. using bis-NHC (N-heterocyclic carbenes) as electron-
donating ligands (52–54). By introducing hydroxyl or sulfonate groups
to the side carbon chains, the water solubility was improved; thus, the
activity was remarkably enhanced for the hydrogenation of CO2 to

CnMen OH CnMen
N N N N
HO
M Cl M
Cl Cl N PF6
N SO3 I Ir I O3 S
N N
N N K K
N N O
O

9 M: Ru; n = 6 11 M: Ru; n = 6
10 M: Ir; n = 5 12 M: Ir; n = 5 13
Figure 6.2 Peris's NHC complexes for CO2 hydrogenation in water.
198 Wan-Hui Wang et al.

0
-H+ -H+
OH2 OH2 OH
Ir Ir Ir
N +H+ N +H+ N
N CO2H N CO2 N CO2

pKa1: 4.0 pKa2: 9.5


14 14ⴕ 14ⴖ
Figure 6.3 Fukuzumi's catalyst for CO2 hydrogenation in water.

HCO2K. Finally, a high TON of 190,000 was achieved with complex


IrI2(AcO)(bis-NHC), 13, under 6 MPa H2/CO2 (1/1) at 200  C in 75 h.
Recently, Fukuzumi et al. developed a C,N-chelated water-soluble Ir
complex 14 bearing a carboxyl group (55). This complex can deprotonate
to give benzoate complex 140 and hydroxo complex 1400 with pKa of 4.0 and
9.5, respectively (Figure 6.3). It was utilized for CO2 hydrogenation in
0.1 M K2CO3 solution by bubbling H2/CO2 (1/1, 50 mL min1) under
atmospheric pressure at 25  C. A TOF of 6.8 h1 and a TON exceeding
100 were obtained over 15 h.

2.2. Design and synthesis of complexes with proton-


responsive ligands
Although great successes in CO2 reduction have been achieved in numerous
previous studies, several critical problems for practical applications remain to
be solved: (1) obtaining highly active catalysts under mild conditions essen-
tial for lowering the overall energy barrier for the conversion of thermody-
namically stable CO2, avoiding the harsh conditions that increase the energy
cost; (2) increasing the cost efficiency of the catalytic process through the
reusability of the catalyst, which usually contains noble metals; and (3)
preventing waste in the catalytic transformations though the avoidance of
volatile organic solvents and additives and the incorporation of more envi-
ronmentally friendly processes. Toward these objectives, it is highly desir-
able to develop effective catalysts that operate in aqueous solution under
mild conditions.
In contrast to the widely used phosphine complexes, N,N-chelated com-
plexes have been less studied in the context of CO2 hydrogenation (60–63).
We observed CO2/H2 generation in the transfer hydrogenation of ketones
with [Cp*Rh(bpy)Cl]Cl in aqueous solutions of FA (64). This result indi-
cated that the Rh complex could catalyze HCO2H dehydrogenation, which
Ligand Design for Effective Catalysis 199

OH OH OH 2

L N L N L N
Cl Cl
M M M
N Cl N H2O N SO42
Cl

OH OH OH
15 M = Ir, L = Cp* 18 M = Ir, L = Cp* 21 M = Ir, L = Cp*
16 M = Rh, L = Cp* 19 M = Rh, L = Cp* 22 M = Rh, L = Cp*
17 M = Ru, L = C6Me6 20 M = Ru, L = C6Me6 23 M = Ru, L = C6Me6

HO 2 HO OH HO N OH 2

Cp* N Cp* N N Cl 2Cl Cp* N


Ir Ir Ir Ir
SO42 SO42
H2O N Cl N N H2O N
Cp*
HO HO OH HO N OH

24 25 26

Figure 6.4 Half-sandwich complexes bearing proton-responsive ligands with hydroxyl


substituents.

is the reverse reaction of CO2 hydrogenation. This prompted us to inves-


tigate CO2 hydrogenation with [Cp*Rh(bpy)Cl]Cl in water. In 2004,
our preliminary studies using [Cp*Rh(bpy)Cl]Cl as a prototype catalyst
showed that this catalyst could hydrogenate CO2 in water albeit at a low rate
(65). Jessop’s experimental results (66) and Sakaki’s theoretical calculations
(67) have demonstrated that the strong electron-donating ability of the
ligand leads to high activity of the complexes. Inspired by these studies,
we designed and synthesized a series of half-sandwich complexes 15–23
(Figure 6.4) by introducing electron-donating hydroxyl groups to the
bpy and phen (1,10-phenathroline) ligands. It is noteworthy that the reac-
tivity of complexes 15–17 is essentially identical to their aqua analogs 21–23
in aqueous solution because of the rapid aquation of the Cl ligand. By
changing the solution pH, the dihydroxyl-substituted ligands 4DHBP and
DHPT (4,40 -dihydroxyl-2,20 -bipyridine, 4,7-dihydroxyl-1,10-phenathroline,
respectively, Figure 6.4) can (de)protonate easily. This property makes
them pH-switchable and imparts proton-responsive electron-donating
functionality owing to the importance of the “keto” resonance structure
of the deprotonated oxyanion ligand (Figure 6.5). The electron-donating
ability is characterized by Hammett constants (sp þ ): the more negative the
200 Wan-Hui Wang et al.

n n-2
OH O O n-2

N -2H+ N N

+2H+
N N N

OH O O

Electron-donor
(OH: s p+ = -0.92) (O-: s p+ = -2.30)
and polarity
Moderate Strong
Hydroxyl form Oxyanion form
Figure 6.5 Acid–base equilibrium between hydroxyl and oxyanion forms.

sp þ values, the stronger the ability to donate electrons. Therefore, the


oxyanion (sp þ ¼ 2:30) is a much stronger electron donor than the
corresponding hydroxyl group (sp þ ¼ 0:92). The acid–base equilibrium
between the hydroxyl and oxyanion forms enabled switching of the polar-
ity and electron-donating ability of the ligand, thus affecting the catalytic
activity and water solubility of the complex. Using this concept, we
achieved highly efficient proton-responsive iridium catalysts for the hydro-
genation of CO2 in water through this sophisticated ligand design (56, 65,
68–70). The high activity of these complexes is attributed to the electron-
donating effect of the deprotonated OH at the para position (vide infra). In
addition, catalyst recycling was achieved using complex 18, which has tun-
able water solubility by changing the solution pH (69).
Complexes with cooperating ligands (i.e., ligand–metal bifunctional com-
plexes) have shown unique properties and remarkable catalytic activity (10–13,
49, 71, 72) Recently, a 2-hydroxylpyridine moiety was identified in the active
site of [Fe]-hydrogenase (73). Theoretical calculations have demonstrated the
important effect of the hydroxyl group in hydrogen activation by [Fe]-
hydrogenase (74). A Rh model complex bearing 2-hydroxylpyridine moieties
has shown better dehydrogenative activity than that without a hydroxyl group
(75). These studies and [Fe–Fe] hydrogenase model studies carried out by
DuBois’ group (13) encouraged us to modify our complex by changing the
OH from the para to the ortho position (57). Thus, the deprotonated OH adja-
cent to the metal center in the complex 24, [Cp*Ir(6DHBP)(H2O)]SO4
(6DHBP: 6,60 -dihydroxyl-2,20 -bipyridine) can be expected to act as both an
electron donor and a pendent base to promote dihydrogen heterolysis. The
activity of the complex 24 was improved markedly over that of the complex
21, as shown in Table 6.1. To introduce more electron donors (while retaining
the pendent bases) and increase the number of active sites, the bipyrimidine-
Ligand Design for Effective Catalysis 201

bridged dinuclear Ir complex 25, [(Cp*Ir)2(THBPM)Cl2]Cl2 (THBPM:


4,40 ,6,60 -tetrahydroxyl-2,20 -bipyrimidine), containing four hydroxyl groups,
was further prepared (58). The catalytic activity of the complex 25 exhibited
a tenfold increase compared with the complex 21 (Table 6.1). It is noteworthy
that the complex 25 was demonstrated to be the most efficient catalyst to date
for CO2 hydrogenation and HCO2H dehydrogenation under mild conditions.
We present the detailed catalytic performance and activation mechanism in the
following sections.

2.3. Mechanism of catalyst activation


2.3.1 Proton-responsive property
To illustrate the properties of the proton-responsive complexes, we carried
out UV–vis titration experiments to determine the pKa values of the com-
plexes 21, 24, and 25. Changing the pH of the solution showed a significant
effect upon the UV–vis absorption. For complex 21, as shown in the chart of
absorbance change versus pH, two pKa values were observed (Figure 6.6).
The first one (pKa1 ¼ 5.0) was attributed to the deprotonation of the two
OH, which have almost the same pKa. The second one (pKa2 ¼ 9.1) was
proposed to be due to the deprotonation of the aqua ligand (Ir–
OH2 ! Ir–OH þ Hþ), assuming that it remains coordinated to the metal
center at pH > pKa1. With the same method, the pKa values of the 6DHBP

Figure 6.6 UV–pH titration with complex 21. Selected single-wavelength data and
Boltzmann fits used to determine the pKa values of the ligand hydroxyl groups and
the H2O ligand (or more loosely bound H2O molecule).
202 Wan-Hui Wang et al.

complex 24 and the THBPM complex 25 were determined to be 4.1 and


3.8, respectively. These results suggest that complexes 21, 24, and 25 are
completely deprotonated in the generally used 1 M NaHCO3 (pH ¼ 8.6)
buffer in the hydrogenation of CO2. It is not clear from these data whether
the pKa2 values correspond to the deprotonation of a water molecule coor-
dinated to the metal center or bound more loosely (e.g., hydrogen bonded
to an oxyanion or N-atom) to the complex. DFT calculations at the
B3LYP/CEP-121g [Ir], 6-311þþG(d,p) [C,N,O,H], 5d//B3LYP/
CEP-121g [Ir], and 6-31þþG(d,p) [C,N,O,H], 5d level of theory with
single-point solvation using the CPCM solvation model with UAHF radii
indicate that complex 24 has a vacant coordination site through loss of the
aqua ligand when the two hydroxyl groups are deprotonated (vide infra) (57).

2.3.2 Electronic effect


When the electron-donating hydroxyl groups were introduced to the bpy or
phen ligands, significantly improved catalytic activity was achieved for the
hydrogenation of CO2. Table 6.2 compares the productivity of the half-
sandwich catalyst [(CnMen)M(L)Cl]þ (M ¼ Rh, Ir, n ¼ 5; M ¼ Ru, n ¼ 6)
and the hydroxyl-substituted analogs (56). The TONs of the iridium cata-
lysts with hydroxyl groups were 50–100 times greater than those of the
unsubstituted catalysts. The electronic effect of substituents was systemati-
cally investigated using [(CnMen)M(4,40 -R2-bpy)Cl]þ (M ¼ Ir, Rh, Ru;
R ¼ OH, OMe, Me, H). As shown in Figure 6.7, the Hammett plots show
a good correlation between the initial TOFs and the sp þ values for three
kinds of complexes (76). The initial TOF (5100 h1) of 15 is over 1000
times higher than that of the unsubstituted analog (4.7 h1) under the same
conditions (1 MPa, 80  C). To our knowledge, it is the first example of cat-
alyst activation by formation of an oxyanion. On the other hand, the

Table 6.2 Substituent effect of ligand on TON for hydrogenation of CO2a


TON
Catalyst L: bpy 4DHBP phen DHPT
[Cp*Rh(L)Cl]Cl 216 b
1800 220 2300
b
[(C6Me6)Ru(L)Cl]Cl 68 4400 78 5100
[Cp*Ir(L)Cl]Cl 105 b
5500 59 6100
a
The reaction was carried out with a cat. (0.1 mM) in a 1 M KOH solution under 1 to 4 MPa (CO2/
H2 ¼ 1:1) at 80  C for 20 h.
b
[Cat.] ¼ 0.2 mM.
Ligand Design for Effective Catalysis 203

Figure 6.7 Correlation between initial TOFs and sp þ values of substituents (R) for the
hydrogenation of CO2 catalyzed by [(CnMen)M(4,40 -R2-bpy)Cl]Cl. (a) M ¼ Ir, n ¼ 5; (b)
M ¼ Rh, n ¼ 5; (c) M ¼ Ru, n ¼ 6; R ¼ OH, OMe, Me, H. The reactions were carried out in
an aqueous 1 M KOH solution under 1 MPa (CO2:H2 ¼ 1:1) at 80  C for 20 h (76).

substituent effects on the rhodium and ruthenium complexes were moderate


compared to that of the iridium complex (Figure 6.7, a vs. b and c). It is appar-
ent that the remarkable activity of the 4DHBP catalyst can be attributed to
the strong electron-donating ability of the oxyanion. The maximum cata-
lytic activity (TOF ¼ 42,000 h1, TON ¼ 190,000) of the Ir(4DHBP) cat-
alyst was obtained at 6 MPa and 120  C. Furthermore, the reaction could
proceed at atmospheric pressure. The Ir(DHPT) complex 18 is the first cat-
alyst that was demonstrated to catalyze CO2 hydrogenation near ambient
conditions (30  C, 1 atm H2/CO2). This result suggested that the
corresponding Ir hydride complex appears to be easily generated as a key
intermediate at atmospheric pressure.

2.3.3 Pendent-base effect


When the hydroxyl group was changed from the para to the ortho position,
the catalytic activity for the hydrogenation of CO2 was significantly
improved. It was found that the 6DHBP complex 24 (TOF: 8050 h1)
showed much higher activity than the 4DHBP complex 21 (TOF:
5100 h1) under the same conditions (57). The electronic effect of the
204 Wan-Hui Wang et al.

Figure 6.8 Correlation between initial TOFs and sp þ values of substituents (R) for the
hydrogenation of CO2 catalyzed by (a) [Cp*Ir(4,40 -R2-bpy)(H2O)]SO4 (black diamonds)
and (b) [Cp*Ir(6,60 -R2-bpy)(H2O)]SO4 (R ¼ OH, OMe, Me, H; circles). Reaction conditions:
1 MPa of H2/CO2 (1/1), 80  C, (a) 0.02–0.2 mM catalyst in 1 M KOH; and (b) 0.01–0.2 mM
catalyst in 1 M NaHCO3 (57).

substituents at the 6,60 positions was also investigated using a series of com-
plexes [Cp*Ir(6,60 -R2-bpy)(H2O)]SO4 (R ¼ OH, OMe, Me, H) with dif-
ferent substituents. As shown in the Hammett plots (Figure 6.8B), similar to
the 4,40 -substituted analogs, stronger electron-donating substituents lead to
markedly enhanced reaction rates. Apparently, there is an additional rate
enhancement for complex 24 compared to that of 21 (Figure 6.8). Since
the electron-donating ability of the hydroxyl group at the para and ortho
position should be almost the same, we proposed that the improved rate
is due to the proximity of the hydroxyl group to the metal center. DFT cal-
culations indicate that the adjacent hydroxyl groups deprotonate to give
oxyanions, which act as pendent bases and assist the heterolysis of H2
(Figure 6.9A–D) (57). NMR experiments suggested that 24 converts faster
than 21 to the Ir–H species (D). For example, 95% of 24 converted to the
Ir–H complex after 30 min under 0.2 MPa H2, while only 90% of 21 trans-
formed to the Ir–H complex after 40 h under 0.5 MPa H2. DFT calculations
on complex 24 under basic conditions support our proposed pendent-base
effect and also suggest that CO2 insertion into the Ir–H bond is stabilized by
a weak hydrogen bond between the hydrido ligand and deprotonated
Ligand Design for Effective Catalysis 205

O O O H
HCO2 C
N +H2O N N
OH2 O O
Ir Ir Ir
-H2O
N Cp* N Cp* N Cp*

O O O
A* 34.8 H2 A 0.0 F 17.0

O O
O
C
H N H
N O
H Ir
Ir
N Cp*
N Cp*
O
O
E 25.7
B 52.5 O H O

N H +H+ N H CO2
Ir Ir
+
-H
N Cp* N Cp*

O O
C 9.9 D 0.6
Figure 6.9 Proposed mechanism of CO2 hydrogenation with complex 24. The square in
(A) indicates a vacant coordination site. Computed free energies at pH 8.3 are indicated
in units of kJ mol1 relative to 1 M (A) in aqueous solution and 1 atm H2/CO2 gases. The
calculated change in free energy for the net reaction around the cycle (i.e.,
H2ðgÞ þ CO2ðgÞ ! Hþ ðaqÞ þ HCO2  ðaqÞ ) is 42.0 kJ mol1 at pH 8.3 (57).

pendent base (Figure 6.9E) (57). In addition, the calculations indicate that
heterolysis of dihydrogen is the rate-determining step, not CO2 insertion
as Ogo and Fukuzumi have suggested (77).
Furthermore, we recently found clear evidence from combined exper-
imental and computational studies of the involvement of a water molecule
in the rate-determining heterolysis of H2, and the enhancement of
proton transfer through the formation of a water bridge in CO2 hydroge-
nation catalyzed by bioinspired complexes bearing a pendent base (59).
Table 6.3 shows the kinetic isotope effects in the hydrogenation of CO2
by catalysts 21, 24, and 26, that is, [Cp*Ir(TH4BPM)(H2O)]SO4
(TH4BPM ¼ 2,20 ,6,60 -tetrahydroxy-4,40 -bipyrimidine).
As shown in Table 6.3, when D2/CO2 (entry 3) was used instead of
H2/CO2 (entry 1), the rate of the reaction in an aqueous solution of
206 Wan-Hui Wang et al.

Table 6.3 Kinetic isotope effect in the hydrogenation of CO2 catalyzed by Ir complexesa
Complex Complex Complex
21b 24b 26b
Gas (1/1, Reaction TOF KIEc TOF KIEc TOF KIEc
Entry 1 MPa) sol. (2 M) Product (h1) (h1) (h1)
1 H2/CO2 KHCO3/ HCO2K 683 – 1730 – 2730 –
H2O
2 H2/CO2 KDCO3/ HCO2K 697 0.98 1520 1.14 1800 1.51
D2O
3 D2/CO2 KHCO3/ DCO2K 572 1.19 1610 1.07 1990 1.37
H2O
4 D2/CO2 KDCO3/ DCO2K 570 1.20 1360 1.27 1190 2.29
D2O
a
The reaction was carried out with catalyst (0.2 mmol) in 10 mL bicarbonate solution (2 M) under
1 MPa H2/CO2 or D2/CO2 (1/1) at 50  C for 1 h.
b
Errors of TOFs and KIEs are typically less than 2% and 4%, respectively.
c
KIE ¼ TOF(entry 1)/TOF(entry n), (n ¼ 2, 3, and 4).

KHCO3 decreased markedly (KIE ¼ 1.19, 1.07, 1.37 for complexes 21, 24,
and 26, respectively; Table 6.3, entry 3). The observed KIEs represent a
composite of the individual isotope effects for both heterolysis of D2 to form
Ir–D and CO2 insertion into Ir–D to generate Ir–ODCO. Therefore, it is
not possible to identify the rate-determining step by analysis of the KIE on
the TOF alone. Since the rate-determining step has been previously
predicted by DFT calculations to be the heterolysis of H2, we expect the
heterolysis of D2 to contribute largely to the observed KIE. Interestingly,
we found a special effect of D2O for different complexes when D2O was
used in the reaction with H2/CO2 or D2/CO2. Heavy water has almost
no effect on the reaction rate when 4DHBP complex 21 is used (entries
1 vs. 2 and entries 3 vs. 4 for complex 21). In contrast, a remarkable influence
on the reaction rate is observed when complexes 24 and 26 are used. The
reaction is significantly inhibited upon replacing H2O with D2O whether
using H2/CO2 or D2/CO2 (entries 1 vs. 2 and entries 3 vs. 4 for complexes
24 and 26, respectively). In addition, the reaction rate decreased markedly
with an increase in the D fraction in the reaction solution. When the reac-
tion was carried out with D2/CO2 in KDCO3/D2O, we obtained the
lowest TOF, which indicates the dual effect of deuterated gas and solvent.
Comparing the KIE data for 21 with those for 24 and 26, we can conclude
that heavy water is involved in the rate-limiting heterolysis of dihydrogen
for complex 24 and 26 but not for 4DHBP complex 21.
Ligand Design for Effective Catalysis 207

The participation of H2O in the transition state is further demonstrated


by DFT calculations. Using the deprotonated Cp*Ir(6DHBP)0 as a proto-
type, we identified three different transition states and pathways to two reac-
tion intermediates resulting from the reaction of Cp*Ir(6DHBP)0 and H2
(Equations 6.8–6.10) (59):

Cp Irð6DHBPÞ0  ðH2 OÞ þ H2 ! Cp IrðHÞ2 ðOH2 Þð6DHBPÞ0


ðwith bridging waterÞ
! Cp IrðHÞð6DHBPHÞ þ H2 O ð6:8Þ
Cp Irð6DHBPÞ þ H2 ! Cp IrðHÞ2 ð6DHBPÞ ðasymmetric
0 0

Ir  H bondsÞ ! Cp IrðHÞð6DHBPHÞ ð6:9Þ


Cp Irð6DHBPÞ0 þ H2 ! Cp Irð6DHBPÞðHÞ2 ðsymmetric
Ir  H bondsÞ ! Cp Irð6DHBPÞðH2 Þ ð6:10Þ

Here, 6DHBPH indicates 6-oxyl-60 -hydroxyl-2,20 -bipyridine. The cal-


culations predicted the water-assisted heterolysis pathway shown in Equa-
tion (6.8) to have the lowest activation free energy (relative to
Cp*Ir(6DHBP)0, H2 and H2O(liq)) of 40.3 kJ mol1 (54.6 and 59.3 kJ mol1
for Equations 6.9 and 6.10, respectively). This result supports the hypothesis
that solvent water is involved in the rate-determining step of H2 heterolysis.
In conclusion, our previous DFT studies have demonstrated that the
rate-determining step in the CO2 hydrogenation using our catalysts with
proton-responsive ligands is the heterolysis of H2 (57). The TOF for formate
generation by CO2 hydrogenation was significantly enhanced by a pendent
base through a second-coordination-sphere effect. As our KIE experiments
described earlier suggest, the solvent effect of D2O has a remarkable influ-
ence on the reaction rate most likely through participation in the rate-
determining step (59). Accordingly, we propose that a solvent molecule
(i.e., H2O) may form hydrogen bonds with the pendent base and

PT
0 0 0
O O H H O H H
O
N H2 N PT N H O
H
H
Ir Ir Ir
H 2O H
N N N
Cp* Cp* Cp*
O O O
Scheme 6.2 Proposed mechanism for H2 heterolysis assisted by the pendent base and
a water molecule through a proton relay. The arrows with PT indicate the movement of
protons via a proton relay. The open square indicates a vacant coordination site (59).
208 Wan-Hui Wang et al.

the approaching H2 along the reaction coordinate for heterolysis and partic-
ipate in the proton transfer to the pendent base in the heterolysis of H2
(Scheme 6.2).

2.3.4 Synergistic effect of electron donor and pendent base


Complex 25 combines the electronic and pendent-base effects described
earlier and exhibited outstanding activity for the hydrogenation of CO2.
In 2 M KHCO3, it yielded a high concentration of formate (1.53 M), which
corresponds to the high TON of 153,000 under 4 MPa H2/CO2 (1/1) at
50  C after 34 h. The highest initial TOF of 53,800 h1 was achieved at
5 MPa and 80  C. More importantly, complex 25 could catalyze the reac-
tion under ambient conditions and achieved the highest initial TOF of
70 h1 and highest TON of 7200 in 336 h at 0.1 MPa and 25  C. As shown
in Figure 6.10, the productivity and activity of 25 were strikingly improved
compared to the 4DHBP Ir catalyst 21 (TOF: 7 h1) and 6DHBP Ir catalyst
24 (TOF: 27 h1). The extraordinary and pH-switchable catalytic activity is
attributed to the polyhydroxyl-substituted ligand, which acts as both a pro-
ton relay and strong electron donor.

Figure 6.10 Time course of CO2 hydrogenation in 1 M NaHCO3 under 0.1 MPa H2/CO2
(1:1) at 25  C using (a) 25, (b) 24, and (c) 21 (50 mM). The inset shows the early time
region.
Ligand Design for Effective Catalysis 209

Catalyst 26 incorporates the synergistic effect of electron donor and


pendent-base activation while retaining a mononuclear Ir complex. The sim-
ilar initial TOFs observed with catalysts 24 and 26 toward CO2 hydrogenation
indicate a more significant enhancement from electron donation by the ligand
than on the number of the active metal centers as seen in Table 6.1. At 50  C
and 1 MPa of H2/CO2, the mononuclear complex 26 also showed similar
activity (TOF: 3060 h1) to that of the dinuclear complex 24 (TOF:
4200 h1). Moreover, a high turnover of 28,000 and high concentration of
formate product (0.56 M) were obtained after 24 h (59).

2.4. pH-dependent water solubility and catalyst recycling


As discussed earlier, the acid–base equilibrium not only changes the elec-
tronic properties of the complex but also affects its polarity and thus its water
solubility. We examined the iridium concentrations in a formate
solution with IPC–MS at different solution pH (56). The 4DHBP complex
15 showed pH-dependent solubility but considerable water solubility
(1 ppm) even at the lowest point (pH: 7, Figure 6.11). Thus, it is not suit-
able for efficient catalyst recycling by precipitation from an aqueous formate
solution. To decrease the water solubility, replacing 4DHBP with the
DHPT ligand, complex 18 exhibited negligible solubility in a weakly acidic

Figure 6.11 pH-dependent solubility of (a) 18 and (b) 15 in a 1 M aqueous formate solu-
tion (56).
210 Wan-Hui Wang et al.

Figure 6.12 Recycling of proton-responsive catalyst 18 with tunable solubility.

Table 6.4 Catalyst recycling studies for the conversion of CO2 into formate using DHPT
iridium catalyst 18a
Loaded/recovered Leaching Final conc. of Recovery
Cycle cat. (ppm) iridiumb (ppm) formate (M) efficiency (%)
1 9.0 0.11 0.105 –
2 8.4 0.22 0.104 93
3 7.7 0.42 0.103 92
4 7.0 0.61 0.103 91
a
Optimized conditions: DHPT catalyst 18 (2.5 mmol), 6 MPa of H2/CO2 (1:1), 0.1 M KOH solution
(50 mL), 60  C for 2 h.
b
Determined by ICP–MS analysis.

formate solution. The lowest Ir concentration (ca. 100 ppb) was found at
pH 5 (Figure 6.11). Then, recycling of 18 was investigated in batchwise
cycles based on the concept shown in Figure 6.12. When the added
KOH was completely consumed by the progress of the hydrogenation reac-
tion, the solution pH decreased, and the DHPT catalyst spontaneously pre-
cipitated due to its decreased water solubility at the lower pH. Thus, a
heterogeneous system was formed and catalytic action was “turned off,” that
is, the reaction terminated automatically. The precipitated catalyst could be
Ligand Design for Effective Catalysis 211

recovered by simple filtration. The iridium complex remaining in the filtrate


was found to be less than 2% of the catalyst loading (0.11 ppm). The recov-
ered catalyst retained a high catalytic activity through four cycles, as shown
in Table 6.4 (56). The generated formate can be isolated by evaporating the
filtrate. Therefore, three components in the reaction (i.e., catalyst, product,
and solvent) can be easily separated without significant waste. The recyclable
catalyst possesses pH-tunable catalytic activity and reaction-controlled water
solubility. This strategy has recently been employed for green and efficient
catalyst recycling (78, 79). These results suggest that by careful consideration
of reaction profiles, advantages of both homogeneous and heterogeneous
catalysts can be combined in innovative catalytic systems.

3. DEHYDROGENATION OF FORMIC ACID


3.1. Historical background
The dehydrogenation of FA using homogeneous catalysts has been less stud-
ied, although FA has been widely used as a hydrogen donor in transfer
hydrogenation. Recently, the concept of using FA as an H2 carrier has
received renewed attention (58, 80–85). The decomposition of FA can
either liberate CO2/H2 by dehydrogenation (or decarboxylation) or give
CO/H2O by dehydration (or decarbonylation) (Equation 6.11). However,
CO is a poison to the catalyst in fuel cells. Therefore, considerable effort has
been devoted to efficient production of hydrogen under mild reaction con-
ditions and selective generation of CO-free hydrogen (80, 82, 84). Selected
results are summarized in Table 6.5.
Decarbonylation Decarbonylation
CO þ H2 O ƒƒƒƒƒƒƒƒ HCOOH ƒƒƒƒƒƒƒƒ! CO2 þ H2 ð6:11Þ
In 2008, Laurenczy et al. reported a ruthenium catalyst with the water-
soluble phosphine ligand TPPTS in aqueous solutions of
HCO2H/HCO2Na (86, 87). No CO was detected by FTIR analysis (detec-
tion limit of 3 ppm). Constant hydrogen generation with total
TON > 40,000 was achieved by continuous addition of FA. Independently,
Beller et al. investigated a ruthenium–phosphine catalyst for dehydrogena-
tion of a FA/NEt3 azeotropic mixture under mild conditions (88). The
commercially available ruthenium complex [RuCl2(PPh3)3] showed a high
initial TOF of 2700 h1 (20 min) and a TON of 890 (2 h) at 40  C.
Improvement of catalytic stability was achieved by using [(C6H6)
RuCl2]/DPPE (DPPE: 1,2-bis(diphenylphosphino)ethane). Under
Table 6.5 Dehydrogenation of formic acid or formate
Catalyst Substrate Solvent T ( C) TON TOFa (h1) COb (ppm) References
RuCl3/TPPTS HCO2H/HCO2Na H2 O 120 >40,000 670 n.d. (86, 87)
[Cp*Rh(bpy)Cl]Cl HCO2H/HCO2Na H2 O 40 – 240 – (65)
[Cp*Ir(4DHBP)(H2O)]SO4 HCO2H H2O 90 10,000 14,000 n.d. (95)
HCO2H H2O 40–80 100,000 <13,500 n.d. (95)
[Cp*Rh(4DHBP)(H2O)]SO4 HCO2H/HCO2Na H2 O 80 83,000 7700 n.d. (76)
[Cp*Ir(H2O)(bpm)Ru(bpy)2](SO4)2 HCO2H/HCO2Na H2 O rt 140 420 n.d. (94)
[(Cp*Ir)2(THBPM)Cl2]Cl2 HCO2H/HCO2Na H2 O 80 308,000 158,000 n.d. (58)
HCO2H/HCO2Na H2 O 90 228,000 165,000 n.d. (58)
a
Initial TOF.
b
n.d., not detected.
Ligand Design for Effective Catalysis 213

continuous-flow conditions, a high TON of 260,000 was observed at room


temperature (89). The generated hydrogen was used to drive an H2/O2
PEM fuel cell, which provided a maximum electric power of approximately
47 mW at a potential of 374 mV for more than 29 h (88). They also studied
the iron-based catalyst for the dehydrogenation of FA (90, 91). Wills et al.
studied dehydrogenation of an FA/NEt3 mixture with ruthenium com-
plexes. Although CO (190–440 ppm) was detected by GC, a high TOF
of 18,000 h1 was observed at 120  C (92). Gas production rates as high
as 1.5 L min1 and total gas production of 462 L were obtained in the
continuous-flow manner (93). Fukuzumi reported dehydrogenation of
FA with a heterodinuclear iridium–ruthenium complex [Cp*Ir(H2O)
(bpm)Ru(bpy)2](SO4)2 (bpm: 2,20 -bipyrimidine), which gave an initial
TOF of 426 h1 and TON of 140 in 20 min at room temperature (94).

3.2. pH-dependent activity


Since most complexes that catalyze CO2 hydrogenation can also promote
the reverse reaction, we studied FA dehydrogenation with our complexes
with proton-responsive ligands (58, 76, 95). The reaction was carried out
in an aqueous FA solution or a mixed HCO2H/HCO2Na solution at
60–90  C. Note that no amine additives or organic solvents were required.
We observed evolution of H2/CO2 from aqueous FA solution with a TOF
of 240 h1 using [Cp*Rh(bpy)Cl]Cl at 40  C (65). Remarkable improve-
ment of catalytic activity was achieved using iridium catalyst [Cp*Ir
(4DHBP)(H2O)](SO4) (21), which is activated by the electronic effect of
the OH groups (95). A high initial TOF of 14,000 h1 and TON of
100,000 was achieved using Ir complex 21 at 90  C in a 2 M aqueous solu-
tion of FA. We also observed that the iridium complex exhibited higher cat-
alytic activity than the rhodium and ruthenium analogs. The dinuclear
complex 25 based on the THBPM ligand showed unprecedented activity,
achieving a TON (308,000, at 80  C) and a TOF (228,000 h1, at 90  C)
higher than ever reported (58). This catalytic system can be utilized in the
reversible storage of H2 by controlling the reaction direction (i.e., forma-
tion/dehydrogenation of FA) through adjusting the pH of the reaction solu-
tion (vide infra).
A study of the pH dependence indicated that the tendency of Ir complex
21 is different from that of the Rh analog 22 and THBPM complex 25
(Figure 6.13). Using 21, the maximum catalytic activity was obtained in
1 M HCO2H (pH 1.8). An increase in pH by the addition of formate led
214 Wan-Hui Wang et al.

Figure 6.13 pH-dependence of the reaction rate using (a) 25 (0.2 mM, closed squares),
(b) 21 (0.2 mM, open diamonds), and (c) 22 (0.2–0.4 mM, closed circles) at 60  C in a 1 M
HCO2H/HCO2Na solution (10 mL).

to decreased initial TOFs. In contrast, using 22 and 25, maximum TOFs


were obtained at pH 2.5 and 3.8, respectively. Complete consumption of
FA was observed in 1 M HCO2H in all cases. In the case of mixed solutions
of FA and sodium formate, only partial consumption of formate was gener-
ally observed. The reaction was significantly inhibited at higher pH (pH > 5)
and negligible gas evolution was detected in a sodium formate solution
(pH 7.6). However, complex 25 can completely convert all the formate
albeit at a low rate.

3.3. Electronic effect for catalyst activation


The electronic effect of the hydroxyl group was also studied in the context of
FA dehydrogenation in acidic solutions (76, 95). The results indicated that
the initial TOF values of iridium and ruthenium complexes correlated well
with the Hammett constants of the substituents on the ligands (Figure 6.14).
Note that the hydroxyl-substituted complexes exist in their protonated
forms in acidic solution. The TOF of 21 with a hydroxyl group
(sp þ ¼ 0:91) was about 90 times that of the unsubstituted analog,
Ligand Design for Effective Catalysis 215

Figure 6.14 Hammett plot of the initial TOF versus sp þ value of the substituent (R) for
two series of complexes: (a) [Cp*Ir(4,40 -R2-bpy)(H2O)]SO4 and (b) [(C6Me6)Ru(4,40 -R2-
bpy)Cl]Cl (R ¼ OH, OMe, Me, H, CO2H). The reaction was carried out in the presence
of catalysts (0.5–2.0 mM) at 60  C in 10 mL of either 1 M HCO2H (for Ir complexes) or
HCO2H/HCO2Na solution (for Ru complexes).

[Cp*Ir(H2O)bpy]SO4. In contrast, the TOF of [(C6Me6)Ru(4DHBP)Cl]Cl


(17) was approximately 2.9 times that of the corresponding unsubstituted
catalyst. In addition, the carboxyl-substituted Ru complex showed poor cat-
alytic activity due to the electron-withdrawing effect of the carboxyl group
(sp þ ¼ 0:42).

3.4. Generation of high-pressure H2 for practical use


Using these catalysts, we have succeeded in providing pressurized gases
(H2/CO2), which is a prerequisite for practical applications (95). The H2
and CO2 can be separated if desired. A spontaneous increase of gas pressure
was observed when the reaction was carried out in a closed system. As shown
in Figure 6.15, the pressure reached 4–5 MPa using 21, 22, and 25. We
found no inhibition in the catalytic system since the conversion of FA
was more than 99%.
We also tested hydrogen generation in a high concentration of formate
for a long period with complex 21. In a 50 mL sample of 8 M FA solution
using 4 mmol of complex 21, the dehydrogenation was carried out at
216 Wan-Hui Wang et al.

Figure 6.15 Time course of reaction pressure in an autoclave (a) using 25 (0.5 mmol) in
10 mL of 2 M FA solution at 60  C, (b) using 22 (2 mM) in 10 mL of 2 M HCO2H/HCO2Na
(95:5) at 80  C, and (c) using 21 (1 mmol) in 10 mL of 2 M FA solution at 80  C.

Figure 6.16 Time course of gas evolution (H2/CO2: 1/1) from dehydrogenation of FA
using 21 (4 mmol) in 50 mL FA solution (8 M) at 40–80  C. The values in parenthesis
are average TOF (h1) (95).
Ligand Design for Effective Catalysis 217

40–80  C for 34 h. As shown in Figure 6.16, gas evolution can be controlled


by adjusting the temperature. All the FA was fully decomposed and no deac-
tivation was observed. The TON reached as high as 102,000 and 20 L of
gases was released in 34 h. In addition, no CO was detected by GC (detec-
tion limit of <10 ppm) in the evolved gas using any of the complexes. These
results indicated promising applicability for practical use.

4. REVERSIBLE HYDROGEN STORAGE BY


INTERCONVERSION OF CO2/H2 AND HCO2H
Although great success has been achieved for either CO2 hydrogena-
tion to formate or FA dehydrogenation, efficient catalytic systems for both
reactions are quite limited (39, 55, 96–98). Further research into the prac-
tical applications for reversible hydrogen storage through interconversion
between H2/CO2 and FA is highly desirable (Figure 6.17). However,
30 years after the concept of H2 storage was proposed (99), only a few exam-
ples of reversible H2 storage have been reported either in an organic solvent
or in water (58, 100–103).
In 1994, Leitner and coworkers developed an FA/NEt3/acetone system
catalyzed by in situ catalyst [Rh(cod)(m-Cl)2]/DPPB (cod: 1,5-
cyclooctadiene; DPPB: 1,2-bis(diphenylphosphino)butane) for hydrogen
storage, although only one and a half cycles were reported (100). Absorption
and release of H2/CO2 were controlled by reaction temperature and pres-
sure. Beller and coworkers investigated the hydrogenation of CO2 in
H2O/THF and dehydrogenation of formate in H2O/DMF in the presence
of [RuCl2(C6H6)2]2 and DPPM (101). The rate of production of formate is
unsatisfying: a TOF of 860 h1 at 70  C under 8 MPa of H2/CO2 (5/3).
The selective dehydrogenation of formate using the same catalyst was
achieved with a TOF of 2900 h1 at 60  C without CO generation

Figure 6.17 Reversible hydrogen storage by the interconversion between H2/CO2


and FA.
218 Wan-Hui Wang et al.

(<1 ppm). Combining the two reactions leads to a reversible hydrogen stor-
age system. Joó et al. demonstrated a charge/discharge device for hydrogen
storage and delivery based on the combination of the hydrogenation of
bicarbonate and dehydrogenation of formate in aqueous solution using
[RuCl2(TPPMS)2]2 without organic additives (102). The hydrogenation
at 10 MPa and 83  C gave formate with a yield of 90% in 200 min. Subse-
quently, the dehydrogenation in a closed system, initially at 1 atm, led to the
release of H2 with 40–50% conversion of the formate. The hydrogenation/
dehydrogenation cycle was repeated 3 times. Most recently, Laurenczy and
Beller et al. tested the in situ catalyst [RuCl2(C6H6)2]/DPPE for dehydroge-
nation of FA and obtained a TOF of 47,970 h1 at 80  C in the presence of
N,N-dimethylhexylamine (103). With constant addition of FA, the highest
TON of 800,000 was achieved without substantial deactivation of the cat-
alyst. They also achieved a high FA/NEt3 ratio of 2.31 (TON 3190) for CO2
hydrogenation using the well-defined complex [RuH2(DPPM)2] instead of
the in situ catalyst in DMF. A reversible “hydrogen-battery” system was
developed by [RuH2(DPPM)2]-catalyzed interconversion between CO2/
H2 and FA/NEt3. Addition of an amine additive after each run is required
due to the loss of NEt3 in the H2/CO2 gas evolution process. The system
showed slight deactivation after seven cycles.
The catalytic activities of iridium, rhodium, and ruthenium catalysts with
proton-responsive ligands have been investigated in the hydrogenation of
CO2 at basic conditions and dehydrogenation of FA at acidic conditions (58,
76, 95). The dinuclear THBPM complex 25 has been demonstrated to be

Figure 6.18 Consecutive hydrogenation of CO2 and dehydrogenation of FA sequence.


The hydrogenation was carried out under atmospheric conditions (0.1 MPa, 30  C), and
the dehydrogenation proceeded in a closed system to provide pressurized gas
(2.3 MPa): (1) hydrogenation of CO2/bicarbonate, (2) acidification by addition of sulfuric
acid, (3) dehydrogenation of FA, and (4) addition of solid KHCO3 to start next run.
Ligand Design for Effective Catalysis 219

the most effective catalyst for CO2 hydrogenation and HCO2H dehydroge-
nation under mild conditions. Consequently, a proof-of-concept study by
the combination of hydrogenation of CO2 and dehydrogenation of FA
using the proton-responsive THBPM iridium catalyst 25 was performed
(Figure 6.18). It exhibited unprecedented activity toward both CO2 hydro-
genation (TOF: 70 h1, 25  C, 1 atm H2/CO2) and HCO2H dehydroge-
nation (TOF: 228,000 h1, 80  C). The catalyst provided a high
concentration of formate (0.48 M) by hydrogenation of CO2 at ambient
temperature and pressure in 2 M KHCO3. In a closed system, acidifying
the solution triggered the release of CO-free H2/CO2 gas and finally
reached 2.3 MPa. Only 17 mM HCO2H was detected after the dehydroge-
nation. The addition of bicarbonate to the reaction solution containing cat-
alyst at ambient conditions restarted the hydrogenation. Two cycles were
achieved without substantial deactivation; the first reversible hydrogen stor-
age with the same catalyst was achieved simply by changing the pH of the
aqueous reaction solution under mild conditions.

5. CONCLUDING REMARKS
The interconversion of CO2 and FA under mild conditions in water
has been recently achieved using Ru and Ir complexes and even using non-
precious metal Fe complexes. Both the TOF and TON have been remark-
ably improved. Only a few catalysts seem to have ligands with several
bifunctional properties such as proton-responsive properties, pendent bases
or acids for a second-coordination-sphere interaction, hydrogen-bonding
functions with a solvent molecule or an added reagent, and electro-
responsive properties. The most successful catalyst is [Cp*Ir
(H2O)]2(THBPM)4þ that has the first three of these characteristics associated
with its bridging ligand. By utilizing the acid–base equilibrium for proton
removal, the ligand becomes a strong electron donor resulting in Ir(I) char-
acter with a vacant coordination site at each metal center in basic solution.
Complemented by DFT calculations, the rates of formate production
using the related Ir mononuclear complexes with and without such
functions on the ligand reveal that the rate-determining step for the CO2
hydrogenation is likely to be the facile H2 heterolysis assisted by the
second-coordination-sphere interaction of the pendent base. These Ir com-
plexes with bifunctional ligands show interconversion of CO2 and FA under
mild conditions in water due to the flat reaction profiles for both directions
as can be seen in biological enzymic systems.
220 Wan-Hui Wang et al.

ACKNOWLEDGMENTS
Y. H. and W.-H. W. thank the Japan Science and Technology Agency (JST), ACT-C for
financial support. The work at BNL was carried out under contract DE-AC02-98CH10886
with the US Department of Energy and supported by its Division of Chemical Sciences,
Geosciences, & Biosciences, Office of Basic Energy Sciences.

REFERENCES
1. Schneider, J.; Jia, H. F.; Muckerman, J. T.; Fujita, E. Chem. Soc. Rev. 2012, 41, 2036.
2. Morris, A. J.; Meyer, G. J.; Fujita, E. Acc. Chem. Res. 2009, 42, 1983.
3. Fujita, E.; Chou, M.; Tanaka, K. Appl. Organometal. Chem. 2000, 14, 844.
4. Agarwal, J.; Fujita, E.; Schaefer, H. F., III; Muckerman, J. T. J. Am. Chem. Soc. 2012,
134, 5180.
5. Agarwal, J.; Sanders, B. C.; Fujita, E.; Schaefer, H. F., III; Harrop, T. C.;
Muckerman, J. T. Chem. Commun. 2012, 48, 6797.
6. Schneider, J.; Jia, H.; Kobiro, K.; Cabelli, D. E.; Muckerman, J. T.; Fujita, E. Energy
Environ. Sci. 2012, 5, 9502.
7. Hayashi, Y.; Kita, S.; Brunschwig, B. S.; Fujita, E. J. Am. Chem. Soc. 2003, 125, 11976.
8. Polyansky, D. E.; Cabelli, D.; Muckerman, J. T.; Fukushima, T.; Tanaka, K.; Fujita, E.
Inorg. Chem. 2008, 47, 3958.
9. Crabtree, R. H. New J. Chem. 2011, 35, 18.
10. Milstein, D. Top. Catal. 2010, 53, 915.
11. Conley, B. L.; Pennington-Boggio, M. K.; Boz, E.; Williams, T. J. Chem. Rev. 2010,
110, 2294.
12. Grützmacher, H. Angew. Chem. Int. Ed. 2008, 47, 1814.
13. Rakowski DuBois, M.; DuBois, D. L. Chem. Soc. Rev. 2009, 38, 62.
14. Reda, T.; Plugge, C. M.; Abram, N. J.; Hirst, J. PNAS 2008, 105, 10654.
15. Inoue, Y.; Izumida, H.; Sasaki, Y.; Hashimoto, H. Chem. Lett. 1976, 863.
16. Munshi, P.; Main, A. D.; Linehan, J. C.; Tai, C. C.; Jessop, P. G. J. Am. Chem. Soc.
2002, 124, 7963.
17. Jessop, P. G.; Ikariya, T.; Noyori, R. Nature 1994, 368, 231.
18. Jessop, P. G.; Hsiao, Y.; Ikariya, T.; Noyori, R. J. Am. Chem. Soc. 1996, 118, 344.
19. Leitner, W.; Dinjus, E.; Gassner, F. In Aqueous-Phase Organometallic Catalysis, Concepts
and Applications; Cornils, B.; Herrmann, W. A. Eds.; Wiley-VCH: Weinheim, 1998,
p 486.
20. Jessop, P. G.; Joó, F.; Tai, C.-C. Coord. Chem. Rev. 2004, 248, 2425.
21. Jessop, P. G. In Handbook of Homogeneous Hydrogenation; De Vries, J. G.; Elsevier, C. J.
Eds.; Vol. 1. Wiley-VCH: Weinheim, 2007; p 489.
22. Federsel, C.; Jackstell, R.; Beller, M. Angew. Chem. Int. Ed. 2010, 49, 6254.
23. Wang, W.; Wang, S.; Ma, X.; Gong, J. Chem. Soc. Rev. 2011, 40, 3703.
24. Wang, W.-H.; Himeda, Y. In Hydrogenation, InTech, 2012; p 249.
25. Fujita, E.; Muckerman, J. T.; Himeda, Y. Biochim. Biophys. Acta Bioenerg. 2013, 1827,
1031.
26. Gassner, F.; Leitner, W. J. Chem. Soc. Chem. Commun. 1993, 1465.
27. Kovacs, G.; Schubert, G.; Joó, F.; Papai, I. Catal. Today 2006, 115, 53.
28. Joszai, I.; Joó, F. J. Mol. Catal. A Chem. 2004, 224, 87.
29. Katho, A.; Opre, Z.; Laurenczy, G.; Joó, F. J. Mol. Catal. A Chem. 2003, 204, 143.
30. Elek, J.; Nadasdi, L.; Papp, G.; Laurenczy, G.; Joó, F. Appl. Catal. A Gen. 2003,
255, 59.
31. Joó, F.; Laurenczy, G.; Karady, P.; Elek, J.; Nadasdi, L.; Roulet, R. Appl. Organometal.
Chem. 2000, 14, 857.
Ligand Design for Effective Catalysis 221

32. Joó, F.; Laurenczy, G.; Nadasdi, L.; Elek, J. Chem. Commun. 1999, 971.
33. Laurenczy, G.; Joó, F.; Nadasdi, L. Inorg. Chem. 2000, 39, 5083.
34. Horvath, H.; Laurenczy, G.; Katho, A. J. Organomet. Chem. 2004, 689, 1036.
35. Erlandsson, M.; Landaeta, V. R.; Gonsalvi, L.; Peruzzini, M.; Phillips, A. D.;
Dyson, P. J.; Laurenczy, G. Eur. J. Inorg. Chem. 2008, 2008, 620.
36. Laurenczy, G.; Jedner, S.; Alessio, E.; Dyson, P. J. Inorg. Chem. Commun. 2007, 10, 558.
37. Federsel, C.; Jackstell, R.; Boddien, A.; Laurenczy, G.; Beller, M. ChemSusChem 2010,
3, 1048.
38. Tanaka, R.; Yamashita, M.; Nozaki, K. J. Am. Chem. Soc. 2009, 131, 14168.
39. Tanaka, R.; Yamashita, M.; Chung, L. W.; Morokuma, K.; Nozaki, K. Organometallics
2011, 30, 6742.
40. Ahlquist, M. S. G. J. Mol. Catal. A Chem. 2010, 324, 3.
41. Yang, X. ACS Catal. 2011, 1, 849.
42. Li, J.; Yoshizawa, K. Bull. Chem. Soc. Jpn. 2011, 84, 1039.
43. Suh, H.-W.; Schmeier, T. J.; Hazari, N.; Kemp, R. A.; Takase, M. K. Organometallics
2012, 31, 8225.
44. Schmeier, T. J.; Dobereiner, G. E.; Crabtree, R. H.; Hazari, N. J. Am. Chem. Soc.
2011, 133, 9274.
45. Langer, R.; Diskin-Posner, Y.; Leitus, G.; Shimon, L. J.; Ben-David, Y.; Milstein, D.
Angew. Chem. Int. Ed. 2011, 9948.
46. Vogt, M.; Gargir, M.; Iron, M. A.; Diskin-Posner, Y.; Ben-David, Y.; Milstein, D.
Chemistry 2012, 18, 9194.
47. Huff, C. A.; Kampf, J. W.; Sanford, M. S. Organometallics 2012, 31, 4643.
48. Zhang, J.; Leitus, G.; Ben-David, Y.; Milstein, D. Angew. Chem. Int. Ed. 2006, 45,
1113.
49. Praneeth, V. K. K.; Ringenberg, M. R.; Ward, T. R. Angew. Chem. Int. Ed. 2012, 51,
10228.
50. Gunanathan, C.; Milstein, D. Acc. Chem. Res. 2011, 44, 588.
51. Kang, P.; Cheng, C.; Chen, Z.; Schauer, C. K.; Meyer, T. J.; Brookhart, M. J. Am.
Chem. Soc. 2012, 134, 5500.
52. Sanz, S.; Benı́tez, M.; Peris, E. Organometallics 2010, 29, 275.
53. Sanz, S.; Azua, A.; Peris, E. Dalton Trans. 2010, 39, 6339.
54. Azua, A.; Sanz, S.; Peris, E. Chem. Eur. J. 2011, 17, 3963.
55. Maenaka, Y.; Suenobu, T.; Fukuzumi, S. Energy Environ. Sci. 2012, 5, 7360.
56. Himeda, Y.; Onozawa-Komatsuzaki, N.; Sugihara, H.; Kasuga, K. Organometallics
2007, 26, 702.
57. Wang, W.-H.; Hull, J. F.; Muckerman, J. T.; Fujita, E.; Himeda, Y. Energy Environ. Sci.
2012, 5, 7923.
58. Hull, J. F.; Himeda, Y.; Wang, W.-H.; Hashiguchi, B.; Periana, R.; Szalda, D. J.;
Muckerman, J. T.; Fujita, E. Nat. Chem. 2012, 4, 383.
59. Wang, W.-H.; Muckerman, J. T.; Fujita, E.; Himeda, Y. ACS Catal. 2013, 3, 856.
60. Bolinger, C. M.; Sullivan, B. P.; Conrad, D.; Gilbert, J. A.; Story, N.; Meyer, T. J.
J. Chem. Soc. Chem. Commun. 1985, 796.
61. Lau, C. P.; Chen, Y. Z. J. Mol. Catal. A Chem. 1995, 101, 33.
62. Caix, C.; ChardonNoblat, S.; Deronzier, A. J. Electroanal. Chem. 1997, 434, 163.
63. Hayashi, H.; Ogo, S.; Abura, T.; Fukuzumi, S. J. Am. Chem. Soc. 2003, 125, 14266.
64. Himeda, Y.; Onozawa-Komatsuzaki, N.; Sugihara, H.; Arakawa, H.; Kasuga, K.
J. Mol. Catal. A Chem. 2003, 195, 95.
65. Himeda, Y.; Onozawa-Komatsuzaki, N.; Sugihara, H.; Arakawa, H.; Kasuga, K.
Organometallics 2004, 23, 1480.
66. Tai, C. C.; Pitts, J.; Linehan, J. C.; Main, A. D.; Munshi, P.; Jessop, P. G. Inorg. Chem.
2002, 41, 1606.
222 Wan-Hui Wang et al.

67. Ohnishi, Y. Y.; Matsunaga, T.; Nakao, Y.; Sato, H.; Sakaki, S. J. Am. Chem. Soc. 2005,
127, 4021.
68. Himeda, Y. Eur. J. Inorg. Chem. 2007, 3927.
69. Himeda, Y.; Onozawa-Komatsuzaki, N.; Sugihara, H.; Kasuga, K. J. Am. Chem. Soc.
2005, 127, 13118.
70. Himeda, Y.; Onozawa-Komatsuzaki, N.; Sugihara, H.; Kasuga, K. J. Photochem. Pho-
tobiol. A Chem. 2006, 182, 306.
71. Eisenstein, O.; Crabtree, R. H. New J. Chem. 2013, 37, 21.
72. Lyaskovskyy, V.; de Bruin, B. ACS Catal. 2012, 2, 270.
73. Shima, S.; Lyon, E. J.; Sordel-Klippert, M.; Kauß, M.; Kahnt, J.; Thauer, R. K.;
Steinbach, K.; Xie, X.; Verdier, L.; Griesinger, C. Angew. Chem. Int. Ed. 2004, 43, 2547.
74. Yang, X.; Hall, M. B. J. Am. Chem. Soc. 2009, 131, 10901.
75. Royer, A. M.; Rauchfuss, T. B.; Wilson, S. R. Inorg. Chem. 2008, 47, 395.
76. Himeda, Y.; Miyazawa, S.; Hirose, T. ChemSusChem 2011, 4, 487.
77. Ogo, S.; Kabe, R.; Hayashi, H.; Harada, R.; Fukuzumi, S. Dalton Trans. 2006, 4657.
78. Zuwei, X.; Ning, Z.; Yu, S.; Kunlan, L. Science 2001, 292, 1139.
79. Wang, W.; Zhang, G.; Lang, R.; Xia, C.; Li, F. Green Chem. 2013, 15, 635.
80. Loges, B.; Boddien, A.; Gartner, F.; Junge, H.; Beller, M. Top. Catal. 2010, 53, 902.
81. Joó, F. ChemSusChem 2008, 1, 805.
82. Enthaler, S.; von Langermann, J.; Schmidt, T. Energy Environ. Sci. 2010, 3, 1207.
83. Grasemann, M.; Laurenczy, G. Energy Environ. Sci. 2012, 5, 8171.
84. Johnson, T. C.; Morris, D. J.; Wills, M. Chem. Soc. Rev. 2010, 39, 81.
85. Fukuzumi, S. Eur. J. Inorg. Chem. 2008, 1351.
86. Fellay, C.; Dyson, P. J.; Laurenczy, G. Angew. Chem. Int. Ed. 2008, 47, 3966.
87. Fellay, C.; Yan, N.; Dyson, P. J.; Laurenczy, G. Chem. Eur. J. 2009, 15, 3752.
88. Loges, B.; Boddien, A.; Junge, H.; Beller, M. Angew. Chem. Int. Ed. 2008, 47, 3962.
89. Boddien, A.; Loges, B.; Junge, H.; Gartner, F.; Noyes, J. R.; Beller, M. Adv. Synth.
Catal. 2009, 351, 2517.
90. Boddien, A.; Loges, B.; Gartner, F.; Torborg, C.; Fumino, K.; Junge, H.; Ludwig, R.;
Beller, M. J. Am. Chem. Soc. 2010, 132, 8924.
91. Boddien, A.; Mellmann, D.; Gärtner, F.; Jackstell, R.; Junge, H.; Dyson, P. J.;
Laurenczy, G.; Ludwig, R.; Beller, M. Science 2011, 333, 1733.
92. Morris, D. J.; Clarkson, G. J.; Wills, M. Organometallics 2009, 28, 4133.
93. Majewski, A.; Morris, D. J.; Kendall, K.; Wills, M. ChemSusChem 2010, 3, 431.
94. Fukuzumi, S.; Kobayashi, T.; Suenobu, T. J. Am. Chem. Soc. 2010, 132, 1496.
95. Himeda, Y. Green Chem. 2009, 11, 2018.
96. Gao, Y.; Kuncheria, J. K.; Jenkins, H. A.; Puddephatt, R. J.; Yap, G. P. A. J. Chem. Soc.
Dalton Trans. 2000, 3212.
97. Man, M. L.; Zhou, Z. Y.; Ng, S. M.; Lau, C. P. Dalton Trans. 2003, 3727.
98. Preti, D.; Squarcialupi, S.; Fachinetti, G. Angew. Chem. Int. Ed. 2010, 49, 2581.
99. Williams, R.; Crandall, R. S.; Bloom, A. Appl. Phys. Lett. 1978, 33, 381.
100. Leitner, W.; Dinjus, E.; Gassner, F. J. Organomet. Chem. 1994, 475, 257.
101. Boddien, A.; Gärtner, F.; Federsel, C.; Sponholz, P.; Mellmann, D.; Jackstell, R.;
Junge, H.; Beller, M. Angew. Chem. Int. Ed. 2011, 50, 6411.
102. Papp, G.; Csorba, J.; Laurenczy, G.; Joó, F. Angew. Chem. Int. Ed. 2011, 50, 10433.
103. Boddien, A.; Federsel, C.; Sponholz, P.; Mellmann, D.; Jackstell, R.; Junge, H.;
Laurenczy, G.; Beller, M. Energy Environ. Sci. 2012, 5, 8907.

You might also like