You are on page 1of 36

CHAPTER SEVEN

Catalytic Hydrogenation of
Carbon Dioxide to Formic Acid
Arno Behr, and Kristina Nowakowski
Technical Chemistry, Department of Bio- and Chemical Engineering, Technical University of Dortmund,
Dortmund, Germany

Contents
1. Introduction 224
2. Hydrogenation of Carbon Dioxide 226
2.1 Thermodynamics and mechanism of the formic acid synthesis 228
2.2 Heterogeneously catalyzed hydrogenation 230
2.3 Homogeneously catalyzed hydrogenation 230
3. Continuous Hydrogenation of Carbon Dioxide in Miniplant Scale 247
3.1 Preliminary batch experiments 247
3.2 Miniplant for carbon dioxide hydrogenation 250
3.3 Miniplant experiments 251
4. Conclusions 252
Acknowledgments 253
References 254

Abstract
In recent years, the utilization of carbon dioxide as alternative C1 building block has
gained more and more scientific interest and has been intensely investigated. Especially
the homogeneously catalyzed hydrogenation of carbon dioxide to formic acid and its
derivates has been well studied. Recently, an increase in product formation was
achieved by further development of the homogeneous catalysts. Currently, iridium-
based catalysts offer the highest catalytic activity known in the hydrogenation of carbon
dioxide. The present chapter gives a wide overview of various catalyst systems, which
have been investigated so far. In addition, current research on the continuously oper-
ated hydrogenation of carbon dioxide in miniplant scale with a promising concept for
catalyst recycling is presented.

Keywords: Formic acid, Carbon dioxide, Homogeneous catalysis, Hydrogena-


tion, Catalyst recycling, Miniplant technique

Advances in Inorganic Chemistry, Volume 66 # 2014 Elsevier Inc. 223


ISSN 0898-8838 All rights reserved.
http://dx.doi.org/10.1016/B978-0-12-420221-4.00007-X
224 Arno Behr and Kristina Nowakowski

LIST OF THE USED ABBREVIATIONS


[EMIM]Cl 1-ethyl-3-methylimidazolium chloride
acac acetylacetonate
AcO acetoxy group
bipy 2,20 -bipyridine
cod cyclooctadiene
Cp, Cp* cyclopentadienyl, pentamethylcyclopentadienyl
CYPO Cy2PCH2CH2OCH3
DBF dibutylformamide
DBU 1,8-diazabicyclo[5.4.0]undec-7-ene
dcpb Cy2P(CH2)4PCy2
dcpe 1,2-dicyclohexylphosphinoethane
dhbipy 4,40 -dihydroxy-2,20 -bipyridine
DHphen 4,7-dihydroxy-1,10-phenanthroline
dippe 1,2-(diisopropylphosphino)ethane
DMSO dimethyl sulfoxide
dppb 1,4-bis(diphenylphosphino)butane
dppbts tetrasulfonated 1,4-bis(diphenylphosphino)butane
dppe 1,2-bis(diphenylphosphino)ethane
dppm bis(diphenylphosphino)methane
dppp 1,3-bis(diphenylphosphino)propane
EDTA ethylenediaminetetraacetic acid
EMIM 1-ethyl-3-methylimidazolium
hfacac hexafluoroacetylacetonate
ICP inductively coupled plasma analysis
mTPPMS meta-monosulfonated triphenylphosphine
n.a. not available
NBD bicycle[2.2.1]hepta-2,5-diene
NHC N-heterocyclic carbene
NMR nuclear magnetic resonance
PP3 [P(CH2CH2PPh2)3]
PTA 1,3,5-triaza-7-phosphaadamantane
RT room temperature
sc supercritical
TEtA triethanolamine
THF tetrahydrofuran
TOF turnover frequency
TON turnover number
Tp hydrotris(pyrazolyl)borate
TPPMS monosulfonated triphenylphosphine
TPPTS trisulfonated triphenylphosphine

1. INTRODUCTION
At the present time, the utilization of carbon dioxide (CO2) is of par-
ticular interest. Nature shows us the way: the carbon dioxide existing on
CO2 Chemistry 225

earth’s atmosphere can be fixed very well in organic compounds by photo-


synthesis. Therefore, it is also conceivable to use carbon dioxide as raw
material and carbon source for chemical synthesis.
Carbon dioxide is an alternative and attractive C1 building block com-
pared with other C1-starting materials (1–10). In contrast to usual carbon
sources, such as carbon monoxide (CO) or phosgene (COCl2), carbon diox-
ide is nontoxic, affordable, and available in large quantities as by-product of
various chemical and energetic processes (11–13). On the other hand, car-
bon dioxide is thermodynamically unfavorable, and for the conversion
efficient and highly active catalysts, electrical reductive or photocatalytic
methods (14–22) are required. However, the two last-mentioned methods
are beyond the scope of this chapter. The amount of theoretical investiga-
tions on the homogeneously catalyzed conversion of carbon dioxide is
increasing more and more (23,24).
Carbon dioxide is already being used as a chemical feedstock in the
industrial conversion of some organic compounds. For instance, urea, sali-
cylic acid, and cyclic carbonates are chemical compounds based on carbon
dioxide (25). However, the amount of chemically used carbon dioxide
accounts only for 1% of the total quantity of carbon dioxide, whereby in
2008, the global anthropogenic carbon dioxide emission amounted to
30  109 t per year with a rising trend (12,26).
Another interesting product available via carbon dioxide conversion is
formic acid, a basic chemical that worldwide produced 324,000 t in 2003,
whereby BASF is the largest producer (27). The applications of formic acid
and its salts include a wide variety, for example, starting chemical for esters,
alcohols, or pharmaceutical products. In addition, formic acid is used in the
textile, leather or dye industry and as cleaning or disinfection solution. Cur-
rently, formic acid is commercially produced by the carbonylation of meth-
anol with carbon monoxide to methyl formate and the following hydration
of the ester with water (see Equation 7.1) (28):

ð7:1Þ

Alternative routes for the formic acid production are the following:
• Oxidation of hydrocarbons
• Hydrolysis of formamide
• Carbonylation of a hydroxide (e.g., sodium hydroxide) and subsequent
acid hydrolysis
These processes are generally depending on the toxic carbon monoxide or
synthesis gas. An alternative and economical synthesis route is the
226 Arno Behr and Kristina Nowakowski

hydrogenation of carbon dioxide to formic acid, whereby the usage of car-


bon dioxide offers a one-step reaction control (Equation 7.2):

ð7:2Þ

The back reaction, the decomposition of formic acid to carbon dioxide


and hydrogen with homogeneous or heterogeneous catalysis, is also an
important topic of current research. Consequently, the resulting formic acid
can be used as easily transportable hydrogen storage (29–45). However,
formic acid is not necessarily an effective storage medium for hydrogen
related to the mass of formic acid, but the separation offers advantages: easy
decomposition, carbon dioxide as coupling product, and unproblematic
handling.
This chapter summarizes the studies on the homogeneously catalyzed
hydrogenation of carbon dioxide with various catalysts. In addition, several
separation possibilities and also our own investigations of the continuous
hydrogenation of carbon dioxide in miniplant scale are presented .

2. HYDROGENATION OF CARBON DIOXIDE


The catalytic conversion of carbon dioxide with hydrogen leads to a
wide range of compounds and offers potential access to various basic
chemicals (46–51). However, these conversions of neat CO2 have not
yet been used in industrial scale. Methanol or even carbon monoxide can
be produced with an excess of hydrogen (52–55). The direct hydrogenation
of carbon dioxide to methanol based on the mixture CO/H2/CO2 is already
performed in the industrial production of methanol (27,56). Other possible
hydrogenation products, such as formaldehyde or even carbon monoxide,
can be synthesized by carbon dioxide hydrogenation (see Figure 7.1).
Formic acid derivates, salts, or adducts are formed in the presence of a
third component (Figure 7.2). In the case of the formate esters, an alcohol,
for example, methanol or ethanol, is necessary for giving methyl or ethyl
formate. The yield of formate esters decreases with increasing alkyl chain.
The formation occurs in two steps: the catalytic hydrogenation of carbon
dioxide to formic acid and the subsequent thermal esterification
(Equation 7.3). In most cases, the esterification proceeds in alkaline solution;
nevertheless, the desired reaction is also possible without a base (57,58):
CO2 Chemistry 227

Figure 7.1 Hydrogenation products of carbon dioxide.

Figure 7.2 Formic acid derivates based on carbon dioxide and hydrogen.

ð7:3Þ

Jessop and coworkers studied the hydrogenation of supercritical carbon


dioxide (scCO2) with methanol and showed the subsequent esterification in
supercritical medium (59). The synthesis of dimethylformamide in the pres-
ence of the secondary amine dimethylamine is feasible in the same solvent
(60,61). Earlier, it has been reported that several catalysts, such as
(Ph2PCH2CH2PPh2)2CoH, (Ph3P)2(CO)IrCl, and (Ph3P)3CuCl, achieved
great turnover number (TON) in the formation of dimethylformamide
under mild reaction conditions (62).
In 1975, Inoue et al. reported the first synthesis of formates in the pres-
ence of alcohols, especially ethanol, and tertiary amines catalyzed by
Pd(dppe)2 (63). In the following chapters, the hydrogenation of carbon
dioxide to formates will be discussed in detail.
228 Arno Behr and Kristina Nowakowski

2.1. Thermodynamics and mechanism of the formic acid


synthesis
Since the mid-1970s, the hydrogenation of carbon dioxide to formic acid
has been intensely studied. The conversion of the comparatively unreactive
carbon dioxide to pure formic acid is an endergonic process
(△G0 ¼ 32.8 kJ mol1) and therefore thermodynamically unfavorable. In
the presence of an additive, such as Na2CO3, triethylamine (NEt3), or
ammonia, the reaction becomes exergonic (△G0 ¼ 9.5 kJ mol1) and
takes place spontaneously (see Figure 7.3).
Many recent studies are focused on the mechanism of the homoge-
neously catalyzed conversion of carbon dioxide with hydrogen
(57,65,66). Several possible mechanisms are conceivable, whereby the
insertion of carbon dioxide into the M–H bond is identical for all mecha-
nisms. Carbon dioxide forms a metal formate complex via 2-coordination
and migration of hydrogen. The generally accepted mechanism can be pos-
tulated as shown in Figure 7.4 (4,67).
On the other hand, the insertion can theoretically occur by M–H bond
breaking based on the weak H–CO2 interaction (68,69). The following
elimination of formic acid can proceed in four different ways (70):
• Reductive elimination of formic acid (71)
• Addition of hydrogen to the formate complex (57)
• Direct hydrogenolysis of the M–O bond without previous oxidative
addition to hydrogen (66)
• Hydrolysis of the metal formate (72)
Organic solvents, water, ionic liquids, or directly supercritical carbon diox-
ide can be used as solvents for the conversion of carbon dioxide. One of the
first investigations already demonstrated the promoting effect of water
for the homogeneously catalyzed hydrogenation of carbon dioxide (72).
Small amounts of water can increase the catalytic activity, whereas no

Figure 7.3 Thermodynamics of CO2 hydrogenation (64).


CO2 Chemistry 229

Figure 7.4 General mechanism for the homogeneously catalyzed hydrogenation of car-
bon dioxide to formic acid (4).

distinct clarifications have existed until then. Yin et al. investigated and
proved the positive influence of water with the metal hydride species
TpRu(PPh3)(CH3CN)H (Tp ¼ hydrotris(pyrazolyl)borate) in the catalyzed
hydrogenation (73). By experimental and theoretical studies, the catalytic
cycle of the simultaneous transfer of a hydride and a proton from the inter-
mediate TpRu(PPh3)(H2O)H to carbon dioxide to form formic acid was
successfully proven. On the other hand, the presence of water in the
rhodium-catalyzed hydrogenation in nonprotic solvents and tertiary amine
has an unfavorable influence on the catalytic activity (74). Performing the
reaction in water, the employed ligands are needed to be soluble in water.
For this purpose, various water-soluble ligands with different substituents
have been investigated, such as acidic (SO3H) or basic (NR2) functional
groups (75–77).
Carbon dioxide dissolved in water exhibits the following equilibrium
(Equation 7.4):

ð7:4Þ

In addition to the hydrogenation of carbon dioxide, the reduction of


HCO 3 in aqueous solution is feasible by transition metal complexes
(78–81).
At present, a variety of homogeneous catalysts of group VIII transition
metals indicate a high activity for the synthesis of formic acid and its derivates
based on carbon dioxide. In the following chapters, the first investigations of
carbon dioxide hydrogenation in heterogeneous systems and subsequent the
homogenous catalyst systems are presented.
230 Arno Behr and Kristina Nowakowski

2.2. Heterogeneously catalyzed hydrogenation


In 1935, Farlow and Adkins were the first to discover the synthesis of formic
acid by the hydrogenation of carbon dioxide using the heterogeneous cat-
alyst raney nickel. After 1 h reaction time at 80  C, 55% yield of formate was
obtained in the presence of various amines. If the reaction temperature is
above 100  C, the substituted formamide could form via dehydration of
the formate (82).
Previous works proposed the heterogeneously catalyzed hydrogenation,
ruthenium catalysts immobilized on polystyrene resin or silica, and the use of
activated carbon supported ruthenium (83–85). Ruthenium hydride species
are considerably active in the hydrogenation of carbon dioxide (67,86,87).
Hao et al. showed that the g-Al2O3-supported ruthenium catalyst
(2.0 w%) achieved a maximal TON up to 120 at 80  C. Furthermore, a
possible mechanism of the heterogeneously catalyzed hydrogenation could
be demonstrated (88).
A recent patent of BASF illustrates a process for the heterogeneously
catalyzed hydrogenation of carbon dioxide to formic acid, in which the
use of a titania-supported gold catalyst and tertiary amines are described
(89). Previous investigations of Preti et al. presented the catalytic activity
and high stability also towards carbon monoxide of the used gold catalyst
(90,91).
Great advantages of the heterogeneously catalyzed hydrogenations are
the simple separation of the catalyst and the product phase, the stability
and handling of the catalyst, and the easy reusability. However, the catalytic
activity is limited; therefore, an enhancement of the yield can only be
achieved by using homogeneous catalysis.

2.3. Homogeneously catalyzed hydrogenation


In 1976, Inoue et al. reported the first homogenously catalyzed hydrogena-
tion of carbon dioxide. Group VIII transition metal complexes, such as
RhCl(PPh3)3, Pd(dppe)2, and H3Ir(PPh3)3, combined with triethylamine
and water, were used under mild conditions (72).
Based on these results, a variety of homogenous catalysts, in particular
ruthenium- and rhodium-based systems, have been studied for the hydro-
genation of carbon dioxide. At present, iridium catalysts achieved the
highest known yields in the conversion of carbon dioxide. In the following,
a detailed overview will be provided for the homogeneously catalyzed car-
bon dioxide hydrogenation.
CO2 Chemistry 231

2.3.1 Precious metal catalysts for the hydrogenation of carbon dioxide


First, precious metal homogeneous catalysts based on ruthenium, rhodium,
and iridium will be presented, followed by the nonprecious metal catalysts.
As already mentioned, the first investigation for the hydrogenation of
carbon dioxide was carried out with group VIII transition metal complexes.
In the first experimental studies, the yield of formic acid was very low. By
continuous development of potential catalysts, an increase of the formic acid
yield could be achieved.

2.3.1.1 Ruthenium catalysts in the hydrogenation of carbon dioxide


Table 7.1 gives a comprehensive overview of the ruthenium-catalyzed syn-
theses of formic acid and its derivates based on carbon dioxide.
TpRu(PPh3)(CH3CN)H (Tp ¼ hydrotris(pyrazolyl)borate) forms dif-
ferent intermediates in various alcohols. Ng and Yin have suggested that
the catalytic cycle of the carbon dioxide hydrogenation is the same
with nonacidic alcohols as well as with acidic 2,2,2-trifluoroethanol
(CF3CH2OH) (see Figure 7.5). Judging on the high promoting effect,
the acidic alcohol CF3CH2OH has a high catalytic activity (97).
For thermodynamically reasons, it is conventional to employ organic
bases or bicarbonates as additive by using ruthenium or rhodium catalysts.
However, inorganic bases can also be used. [RuCl(C6Me6)(DHphen)]Cl
obtains a very high TON up to 15,400 in combination with the inorganic
base potassium hydroxide (Table 7.1, entry 3). Due to the addition of
base, the majority of the hydrogenation of carbon dioxide is carried out
in basic solution. Hayashi observed the positive influence of a proton on
the reduction of carbon dioxide under acidic condition (pH range 3–5)
catalyzed by the in situ formed hydride complex [(6-C6Me6)Ru(II)(bpy)
H]2þ (109). Further mechanistic investigations were carried out by
Ogo and Fukuzumi with a special emphasis on the hydrogenation of carbon
dioxide in the presence of water-soluble arene ruthenium and iridium aqua
complexes (110,111). Thai et al. showed that the neutral and cationic
complexes [(6-arene)Ru(2-N,O-L)Cl] and [(6-arene)Ru(2-N,O-L)
(H2O)]þ (LH ¼ 8-hydroxyquinoline) catalyzed the hydrogenation of car-
bon dioxide to formate in aqueous solution (107).
The group around Joó extensively investigated the hydrogenation of car-
bon dioxide in amine-free aqueous reaction solution (78,80,112). The cat-
alytic activity depends on the pH value, whereby in certain cases, the yield
increases at a pH level of 8. Especially [RuCl2(TPPMS)2]2 achieved high
TON in the hydrogenation of NaHCO3 (79). Also, Gao has discovered a
Table 7.1 Ruthenium catalysts used for the hydrogenation of carbon dioxide to formic acid and their derivates
Temp. Pressure CO2/H2 Time TON TOF
Entry Catalyst Solvent Base/additive ( C) (bar) (h) () (h1) Refs.
1 [Ru(Cl2bipy)2(H2O)2] EtOH NEt3 50 30/30 8 5000 625 (58)
[CF3SO3]2
2a RuCl2(PMe3)4 scCO2 NEt3 50 120/80 47 7200 n.a. (87)
3 [RuCl(C6Me6)(DHphen)]Cl H2O KOH 120 30/30 24 15,400 642 (92)
b
4 RuCl(OAc)(PMe3)4 scCO2 NEt3 50 120/70 0.3 31,700 95,000 (93)
5 [RuCl2(mTPPMS)2]2 H2O NaHCO3 80 35/60 0.03 320 9600 (79)
a
6 RuH2(PMe3)4 scCO2 NEt3 50 120/85 1 1400 1400 (86)
7 Ru2(CO)5(dppm)2 Acetone NEt3 RT 38/38 21 2160 103 (94)
8a RuCl3/PPh3 EtOH NEt3 60 60/60 5 200 40 (95)
9 CpRu(CO) Benzene NEt3 120 30/30 45 43 1 (96)
(m-dppm)Mo(CO)2Cp
10 TpRuH(PPh3)(CH3CN) THF NEt3 100 25/25 16 1815 113 (73,97)
11a RuH2(PPh3)4 Benzene NEt3 RT 25/25 20 87 4 (72)
12 RuH2(PPh3)4 Benzene Na2CO3 100 25/25 4 169 42 (98)
13 [(C5H4(CH2)3NMe2)Ru(dppm)] THF None 80 40/40 16 8 0.5 (99)
BF4
14 [RuCl2(CO)2]n H2O, NEt3 80 27/81 0.3 400 1300 (100)
iPrOH
15 RuCl3(aq)/dppbts H2O HNMe2 60 25/25 15 4980 n.a. (101)
16 K[RuCl(EDTA-H)Cl]2H2O H2O None 40 17/3 0.5 n.a. 250 (102)
17 RuCl2(PTA4) H2O NaHCO3 80 0/60 n.a n.a 807 (78)
18 [(Cl2bipy)2Ru(H2O)2] EtOH NEt3 150 30/30 8 5000 n.a. (58)
[CF3SO3]2
19 [(NHC)2RuCl][PF6] H2O KOH 200 20/20 75 23,000 n.a. (103)
20 [(C6Me6)Ru(dhbipy)] H2O KOH 120 30/30 8 13,620 4400 (104)
21 [(dppm)Ru(H)(Cl)] H2O NaHCO3 70 0/50 2 1374 687 (105)
22 [RuCl2(TPPMS)2]2 H2O NaHCO3 25 5/35 2 524 262 (80)
23 [Cp*Ru(DHphen)Cl]Cl H2O KOH 120 30/30 24 15,400 3600 (106)
24c [(6-arene)Ru(2-N,O-L) H2O NEt3 100 50/50 10 400 n.a. (107)
(H2O)]BF4
25 RuH2(PMe3)4 scCO2 NEt3 50 120/80 3 1900 630 (87)
26 RuH2(PPh3)4 DBF TEtA 50 30/30 1 698 698 (108)
a
Water as additive.
b
C6F5OH as additive.
LH ¼ 8-hydroxyquinoline.
c

All abbreviations can be found in “List of the Used Abbreviations.”


234 Arno Behr and Kristina Nowakowski

Figure 7.5 Mechanism of the TpRu(PPh3)(CH3CN)H-catalyzed hydrogenation in non-


acidic alcohols (Tp ¼ hydrotris(pyrazolyl)borate) (97).

critical dependence on the pH value for the formation and decomposition of


formic acid in acetone in the presence of Ru2(CO)5(dppm)2 (94).
Nevertheless, in the absence of a base or an additive, the reaction of car-
bon dioxide occurs with very low TONs and TOFs (Table 7.1, entry 13,
16). By reference to these reaction conditions, the necessity of an additive
is revealed. Particularly, the use of the catalyst RuH2(PPh3)4 shows the obvi-
ous influence of various additives. In comparison to triethylamine, the car-
bonate Na2CO3 reaches the 10-fold higher catalytic activity in terms of
turnover frequency (TOF).
With an increasing excess of PMe3 using RuH2(PMe3)4, the TOF is
reduced from 1300 to 360 (87). Musashi and Sakaki suggested that an excess
of phosphine is not advantageous for the reaction but rather the employed
solvent is important for the dissociation of the phosphine (67). By the appli-
cation of further theoretical investigations, it was found that the carbon
dioxide insertion into the Ru(II)–H bond is the rate-determining step,
whereby other reactions steps, such as isomerization of ruthenium formate
and the metathesis of 1-formate intermediate, succeed faster in the exper-
imental investigations. The use of polar solvents is recommended to over-
come the rate-determining step (113).
CO2 Chemistry 235

The bimetallic complexes of Ru–Mo and Ru–W show low catalytic


activity in the formation of formic acid, which is due to the lacking forma-
tion of the required hydride intermediate (96).
Instead of organic solvents, a conceivable alternative for the hydrogena-
tion of carbon dioxide is supercritical carbon dioxide (scCO2) (114,115).
Jessop et al. described the first successful use of supercritical solvents for
the production of formic acid. The desired reaction occurs in a supercritical
mixture of CO2 and H2 in the presence of ruthenium(II) phosphine com-
plexes. Under these conditions, a TON of 1400 was obtained, whereas
organic solvents provided a lower activity and a slower reaction (86). In con-
trast to the chlorine-containing ruthenium phosphine complex
RuCl2(PMe3)4, lower TONs are achieved by using RuH2(PMe3)4 (87).
Both theoretical and experimental investigations were carried out by
Baiker and coworkers (116). In addition, Jessop et al. confirmed that an
enhancement of the TON is possible by using scCO2 as solvent. In the
presence of triethylamine, RuCl(OAc)(PMe3)4 achieved the highest
TON in ruthenium systems up to 31,700 (Table 7.1, entry 4) (93). It is
supposed that the higher yield is attributed to the increased miscibility of
hydrogen in scCO2 (117). Mechanistic investigations of the ruthenium
trimethylphosphine-catalyzed hydrogenation, especially with RuCl(OAc)
(PMe3)4, were carried out using high-pressure NMR spectroscopy (118).
The active catalyst in the reaction is an unsaturated, cationic ruthenium
complex [(PMe3)4RuH]þ. Getty et al. suggested that the used base DBU
is involved in the generation of the active catalyst species.
Leitner et al. have studied the continuous-flow hydrogenation of super-
critical carbon dioxide containing the immobilized catalyst [Ru(cod)(met-
hallyl)2] with [EMIM]Cl and [PBu4][TPPMS] as ligand. The biphasic
system contains a mobile (scCO2) and a stationary phase (ionic liquid).
The scCO2 is simultaneously reactant and extractive phase for the pure
formic acid. By a continuous removal of formic acid, the reaction takes place
with a TOF of 1089 at 50  C and 0.5 h of reaction (119,120). Based on these
results, supercritical carbon dioxide seems to be a promising, alternative sol-
vent for carbon dioxide hydrogenation.
A further ionic liquid system with an immobilized ruthenium catalyst
containing water and a tertiary amine was applied by Zhang et al. (121).
In 2010, Yasaka et al. reported the use of ionic liquids based on imidazole
for the hydrogenation of carbon dioxide to formic acid. The product is sta-
bilized by the formate anions of the ionic liquid (122).
236 Arno Behr and Kristina Nowakowski

2.3.1.2 Rhodium catalysts in the hydrogenation of carbon dioxide


The rhodium catalysts, which are described in literature for the hydrogena-
tion of carbon dioxide, are summarized in Table 7.2. In 1981, Yamaji exam-
ined the hydrogenation of carbon dioxide with the Wilkinson complex
RhCl(PPh3)3 and Na2CO3 as additive in benzene (123). After 3 h reaction
time, a TON up to 173 was reached (Table 7.2, entry 6). In addition,
Ezhova et al. have reported that RhCl(PPh3)3 with triethylamine and
PPh3 as added ligands achieved a distinct higher TON by increasing the
P:Rh ratio (124).
The group of Leitner has extensively investigated the hydrogenation of
carbon dioxide using rhodium complexes with bidentate phosphine ligands
(74,125,126). The in situ formed rhodium complexes are active systems for
the direct formation of formic acid. The generated complex [RhCl
(cod)]2/dppb in dimethyl sulfoxide (DMSO) and under mild reaction con-
dition achieved a TOF up to 52 (Table 7.2, entry 12). The steric and elec-
trochemical properties of the phosphorous ligands have a considerable
influence on the catalytic activity. This is particularly the case in the direct
comparison with dppb and dippe with [RhCl(cod)]2. By using dippe as
ligand, the TOF is lower than with dppb (Table 7.2, entry 5). Kinetic mea-
surements of the rhodium–cyclooctadiene complexes have shown that the
neutral rhodium(I) hydride complexes are the active intermediates in the
rhodium-catalyzed hydrogenation (74). Furthermore, Schindler has studied
the kinetic of the hydrogenation of carbon dioxide in detail using the formed
complex [(dppp)2RhH] and various solvents (127). The cationic formate
complex [(dppp)2Rh]HCOO is generated in DMSO; therefore, the inser-
tion is only possible in DMSO.
At the same time, Tsai and Nicholas have identified the detailed mech-
anism of the hydrogenation of carbon dioxide catalyzed by [Rh(NBD)
(PMe2Ph)3]BF4 in tetrahydrofuran (THF) (see Figure 7.6). In the presence
of hydrogen, the catalytic active rhodium dihydride complex [H2Rh
(PMe2Ph)3(S)] [S ¼ H2O, THF] is formed. Via kinetic and spectroscopic
investigations, it was experimentally proven that the reductive elimination
step is the limiting factor (65). By theoretical investigations, Hutschka and
Dedieu were able to show the same results (128,129).
Pomelli et al. investigated the elimination step of formic acid in the pres-
ence of rhodium complexes by density functional theory and figured
out that the dissociation of formic acid is thermodynamically preferred in
scCO2 (130).
Table 7.2 Rhodium catalysts used for the hydrogenation of carbon dioxide to formic acid and their derivates
Entry Catalyst Solvent Base Temp. ( C) Pressure CO2/H2 (bar) Time (h) TON () TOF (h1) Refs.
1 Rh(hfacac)(dcpb) DMSO NEt3 25 20/20 5 n.a 1335 (132)
2 RhCl(TPPTS)3 H2O NHMe2 RT 20/20 12 3439 n.a. (131)
3 [RhH(cod)]4/dppb DMSO NEt3 RT 20/20 18 2200 375 (74)
4 RhCl(TPPTS)3 H2O NHMe2 81 20/20 0.5 n.a. 7260 (133)
5 [RhCl(cod)]2/dippe DMSO NEt3 24 40 total 18 205 11 (126)
6 RhCl(PPh3)3 benzene Na2CO3 100 55/60 3 173 58 (123)
7 [Rh(2-CYPO)2]BPh4 MeOH NEt3 55 25/25 7 1000 n.a (134)
8 RhCl3/PPh3 H2O NHMe2 50 10/10 10 2150 215 (135)
a
9 [Rh(NBD)(PMe2Ph)3]BF4 THF None 40 48/48 48 128 3 (65)
10 RhCl(PPh3)3/PPh3 MeOH NEt3 25 40/20 20 2700 125 (124)
11 RhCl3(aq)/dppbts H2O HNMe2 60 25/25 15 3910 n.a. (101)
12 [RhCl(cod)]2/dppb DMSO NEt3 RT 20/20 22 1150 52 (125)
13 RhCl(TPPTS)3 H2O NEt3 23 20/20 n.a. n.a. 1364 (133)
14 [Cp*Rh(DHphen)Cl]Cl H2O KOH 80 20/20 32 2400 270 (106)
b,c
15 Rh(NO)(dcpe) n.a. DBU 50 1.5/1.5 16 106 n.a. (136)
d
16 [RhCl(mTPPMS)3] H2O HCOONa 50 50/50 20 65 n.a. (137)
d
17 [RhCl(mTPPMS)3] H2O CaCO3 50 20/80 24 300 n.a. (138)
a
Water as additive.
b
Rh(III)(NO)(dcpe)Cl2 achieved similar results.
c
Addition of the cocatalyst C6F5OH shows no influence.
d
TON was calculated by the given data.
All abbreviations can be found in “List of the Used Abbreviations.”
238 Arno Behr and Kristina Nowakowski

Figure 7.6 Mechanism of the rhodium-catalyzed carbon dioxide hydrogenation inves-


tigated by Tsai and Nicholas (65).

In addition, efficient water-soluble rhodium phosphine complexes, such


as RhCl(TPPTS)3, showed a high catalytic activity in water (Table 7.2,
entry 2). In the presence of dimethylamine and 12 h reaction time, a
TON of 3439 was reached at room temperature (131).
Joó showed that the formation of free formic acid in aqueous solution is
possible in the presence of CaCO3 (138) or HCOONa by using [RhCl
(mTPPMS)3] (137). In the CaCO3 system, the rhodium catalyst achieved
higher yields of formic acid.
A direct comparison of ruthenium- and rhodium-based catalysts shows
that the ruthenium catalyst RuCl3(aq) is significantly more active than
the rhodium analogous (Table 7.1, entry 15 and Table 7.2, entry 11).
The same applies for the half-sandwich complexes [Cp*M(DHphen)Cl]
Cl (M ¼ ruthenium or rhodium) (Table 7.1, entry 23 and Table 7.2, entry
14). The previous theoretical investigations by Musashi and Sakaki have
suggested that the CO2 insertion into M–H bonds will depend on the metal
and its oxidation level. Therefore, the CO2 insertion into the Rh(I)–H and
Ru(II)–H bonds is associated with no barrier and a moderate activation bar-
rier, while for the insertion into the Rh(III)–H bond, a distinctly higher acti-
vation is necessary (139).
In the majority of reactions, the relatively strong base triethylamine is
being used as additive. Triethylamine is inexpensive, is available in large
CO2 Chemistry 239

amounts, and is obtaining higher yields compared to other bases. Leitner and
coworkers have demonstrated this effect by comparing a variety of bases in
the aqueous carbon dioxide hydrogenation with RhCl(TPPTS)3 (133).
Munshi and Jessop investigated various organic and inorganic Brønsted
and Lewis bases in the hydrogenation of carbon dioxide catalyzed by
RuCl(O2CMe)(PMe3)4 and methanol as additive (93). In this context, it
was discovered that DBU achieved distinctly higher yields in contrast to
triethylamine. In the presence of bases with a weak basic character, no
formic acid was formed. The product is a formic acid–amine adduct,
whereas various ratios of 1:1, 2:1, and even 3:1 are possible. The common
ratios of HCOOH/NEt3 adducts are 1.3:1 till 1.8:1 using ruthenium(II) or
rhodium(I) complexes (140), for example, the RuCl2[PMe3]4 complex leads
to a ratio of 1.7:1 (Table 7.1, entry 2). Among others, the solvent is decisive
for this ratio: in nonprotic solvents, ratios higher than 1 are possible due to
the formation of stable 2:1 HCOOH/NEt3 adducts (141). Adduct ratios of
1.6:1 are possible in scCO2 (87). Wagner has indentified that 3:1 adducts are
the most stable adducts, and a thermal separation is possible for the longer-
chain alkyl amines (142). Furthermore, a direct application of the synthe-
sized formic acid–triethylamine adducts as selective hydrogen-transfer
reagents in organic synthesis is feasible (143). However, the formation of
pure formic acid–triethylamine adducts is realized by the conversion of car-
bon dioxide and hydrogen in triethylamine at 40  C under 120 bar CO2/H2
with [RuCl2(PMe3)4] as catalyst. After 1 h reaction time, the 1.78:1 adduct is
formed. By distillation of the solution, the ratio of the adduct changes into
2.35:1, a stable azeotrope. Before the distillation has been performed, the
catalyst is deactivated (140).

2.3.1.3 Separation of the formic acid–amine adducts


To obtain pure formic acid, the product has to split off the salt and the reac-
tion solution. Among others, the formic acid is isolated from the catalyst,
which in most cases could also catalyze the decomposition of formic acid
depending on the reaction conditions. Consequently, the separation of
the formed salts has economical and ecological reasons due to the recovery
of the expensive catalysts. One possibility for the separation is the addition of
another acid to liberate the formic acid or via base exchange with high-
boiling amines, such as imidazole (100,144). The use of a high-boiling
amines includes a thermal decomposition of the adduct, realizing the amine,
which can be recycled again. The negative aspects of the imidazole applica-
tion are an additional and undesired separation step and the need of further
240 Arno Behr and Kristina Nowakowski

solvents (145,146). An additional alternative is the use of multiphase reac-


tion systems, consisting of easily separable catalyst and product phase. Cer-
tainly, the phase containing formic acid has to be purified by subsequent
processing. Behr et al. showed that in aqueous solutions containing RhCl3
or RuCl3 and the water-soluble ligand dppbts (tetrasulfonated 1,4-bis
(diphenylphosphino)butane), an extraction of the resulting formic acid is
technically feasible by using N,N-dibutylformamide (101). Another process
for the separation of formic acid–amine adducts has been investigated by
Paciello and coworkers of BASF. The longer-chain amine trihexylamine
is used as base for the lipophilic [Ru(H)2(PnBu3)4]-catalyzed reaction in
diols. The formic acid–trihexylamine adduct is not miscible with the free
amine but with the polar diol phase, whereas the catalyst remains in the
amine phase. The resulting solution of the salt and the diol can be easily sep-
arated by distillation under mild conditions (147,148). Other researchers
have reported the formation of noncomplexed formic acid in an aqueous
solution (137,138). However, conventional distillation of the mixture is
not economical since large amounts of energy are required due to the azeo-
trope formation. Several patents of BASF demonstrate the removal of formic
acid from aqueous solution by extraction with secondary formamides, such
as N,N-dibutylformamide (149). This method is, among others, applicable
to quaternary ammonium formates (150). Subsequently, the formic acid can
be isolated from the amine by distillation.

2.3.1.4 Other transition metals in the hydrogenation of carbon dioxide


As can be seen from Table 7.3, also other transition metals, such as palla-
dium, have been investigated in the hydrogenation of carbon dioxide. In
general, it can be concluded that palladium shows a lower catalytic activity
in the hydrogenation of carbon dioxide. For the formation of potassium for-
mate with the catalyst PdCl2, Kudo achieved a TON up to 1580 at 160  C
within 3 h reaction time (Table 7.3, entry 1).
Since many years, ruthenium- and rhodium-based catalysts with phos-
phine ligands achieved the best yields in the hydrogenation of carbon diox-
ide. However, current research demonstrates certain iridium-based systems
with very impressive high TONs, whereby the reaction temperature
(120–150  C) is very high in contrast to the conventional ruthenium and
rhodium systems (25–50  C).
In 2004, the half-sandwich iridium(III)complexes with 4,4-dihydroxy-2,2-
bipyridine (dhbipy) or 4,7-dihydroxy-1,10-phenanthroline (DHphen) ligands
were designed by Himeda et al. (104,106,153–155) (Figure 7.7).
CO2 Chemistry 241

Table 7.3 Other metals for the conversion of carbon dioxide to formic acid and their
derivates
Pressure
Temp. CO2/H2 Time TON TOF
Entry Catalyst Solvent Base ( C) (bar) (h) () (h1) Refs.
1 PdCl2 H2O KOH 160 n.a./110 3 1580 530 (151)
2 PdCl2 H2O KOH 240 40/106 3 340 n.a. (151)
a
3 Pd(dppe)2 Benzene NEt3 110 25/25 20 62 3 (72)
4 Pd(dppe)2 Benzene NaOH RT 24/24 20 17 0.9 (152)
5a PdCl2[P Benzene NEt3 RT 50/50 n.a. 15 n.a. (144)
(C6H5)3]2
a
Water as additive.
All abbreviations can be found in “List of the Used Abbreviations.”

Figure 7.7 Half-sandwich iridium(III)complexes (left: 4,4-dihydroxy-2,2-bipyridine ligand


(dhbipy); right: 4,7-dihydroxy-1,10-phenanthroline ligand (DHphen)) (104,106,153–155).

In other works, Joó proposed that the formed formic acid could exceed
the amount of base. The resulting solution changes its pH value from alka-
line to acidic (138). This given fact is used for the performance of the cat-
ionic Cp*iridium(III)complexes in aqueous solution. Depending on the pH
value, the properties of the dhbipy- or DHphen-containing catalyst change
from being dissolved homogeneously in alkaline solution to being precipi-
tated heterogeneously during the reaction. This effect is explained by the
electronic effect and polarity of the oxyanion caused by the phenolic
hydroxyl group. By the use of these iridium complexes, a high catalytic
activity and the possibility of catalyst separation can be achieved. The advan-
tages of homogenous and heterogeneous catalysis are thus combined.
Through modification of [Cp*Ir(dhbipy)Cl]Cl, the iridium-based catalyst
[Cp*Ir(dhbipy)(H2O)]2þ is obtained, which is applicable for the synthesis
of formic acid as well as for the formic acid decomposition in the absence
of organic additives. The reaction can be controlled by regulating the pH
242 Arno Behr and Kristina Nowakowski

value (156). Moret et al. investigated the influence of pH level and temper-
ature both in the hydrogenation of carbon dioxide and in the decomposition
of formic acid via NMR spectroscopy. Important data for the development
of new catalysts for the CO2–formic acid cycle were identified (157).
Two further iridium-based systems are the [IrI2(AcO)(bis-NHC)] com-
plexes containing heterocyclic carbene ligands (NHC) (see Figure 7.8A
and B). Because of the sulfonate functionalities, the water solubility of the
catalyst is increased. Additionally, the systems are more active in the reduc-
tion of carbon dioxide to formate as in absence of the sulfonate groups. The
complex with the bis-abnormal coordination of the NHC ligand currently
shows the best catalytic activity in the reduction with isopropanol to formate
by transfer hydrogenation (158).
Compared with other iridium-based systems, these catalysts reached sim-
ilar TON in the hydrogenation of carbon dioxide with potassium hydroxide
(Table 7.4, entry 5).
In 2009, the Ir(III)–pincer complex [(PNP)IrH3] containing two
diisopropylphosphanyl substituents was announced by Nozaki et al.
(Figure 7.8C) (159). In aqueous solution and with KOH as additive, the
Ir(III)–pincer complex achieved a remarkable TON up to 3,500,000
and a TOF of 73,000 after a reaction time of 48 h at 120  C (Table 7.4,
entry 4). It has to be mentioned that the basicity has a major influence on
the yield. With weaker bases, such as K2PO4, the yield decreases. Currently,
this iridium catalyst is the most active system for the hydrogenation of
carbon dioxide with an impressive high TON.
The mechanism of the iridium(III) trihydride was investigated by means
of density functional theory by Ahlquist (160). The computational studies

Figure 7.8 Iridium-based catalysts: (A) and (B) [IrI2(AcO)(bis-NHC)] complexes (C) Ir(III)–
pincer complex [(PNP)IrH3] (158,159).
Table 7.4 Iridium complexes used for the hydrogenation of carbon dioxide
Entry Catalyst Solvent Base Temp. ( C) Pressure CO2/H2 (bar) Time (h) TON () TOF (h1) Refs.
1 [Cp*Ir(dhbipy)Cl]Cl H2 O KOH 120 30/30 57 42,000 190,000 (154)
2 [Cp*Ir(DHphen)Cl]Cl H2 O KOH 120 30/30 48 222,000 33,000 (106)
3 [Cp*IrCl(DHphen)]Cl H2 O KOH 120 30/30 10 21,000 2100 (92)
4 [(PNP)IrH3] H2O KOH 120 30/30 48 3,500,000 73,000 (159)
5 [IrI2(AcO)(bis-NHC)] H2 O KOH 200 30/30 75 190,000 n.a. (158)
All abbreviations can be found in “List of the Used Abbreviations.”
244 Arno Behr and Kristina Nowakowski

were simplified by replacing the isopropyl groups through hydrogen. The


formate complex is formed via two steps. Initially, a H-bound formate com-
plex and a subsequent more stable O-bound complex are formed. The
formate complex is replaced by hydrogen coordinating to the iridium(III)
center. It was found that the iridium(I) hydride is highly active and thermo-
dynamically advantageous. Yang complemented this research work and
showed that the analogous catalysts based on iron and cobalt are promising
alternatives (161).
The disadvantages of using precious metals without catalyst recycling
are the high costs of catalyst and the low availability of some metals.
The employment of nonprecious catalysts is an interesting alternative.
The first investigations of iron- and cobalt-based systems are presented in
section 2.3.2.

2.3.2 Nonprecious metal catalysts for the hydrogenation of carbon


dioxide
In addition to the active ruthenium, rhodium, and iridium systems, several
nonprecious homogenous catalysts are known for the hydrogenation of car-
bon dioxide. In previous works, Evans and Newell presented the iron cat-
alyst [HFe3(CO)11] for the synthesis of methyl formate (162). Significantly
higher activities can be achieved by the combination of FeCl3 (Table 7.5,
entry 1) with bidentate phosphine ligands (163). However, for the desired
reaction, the expensive base 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU)
and 100 bar absolute pressure are required.
In addition to the iron-based systems, nickel, titanium, and molybdenum
are investigated in the hydrogenation of carbon dioxide (see Table 7.5). As
an example, NiCl2(dcpe) reaches a comparable TON to ruthenium- or
rhodium-based systems, however only after 216 h of reaction (Table 7.5,
entry 5).
Since 2010, the group of Beller has been investigating new nonnoble
homogeneous catalyst systems based on iron and cobalt. The iron catalyst
Fe(BF4)26H2O can reduce carbon dioxide and even bicarbonates. Com-
bined with the phosphorous ligand [P(CH2CH2PPh2)3], abbreviated as
PP3, the system Fe(BF4)26H2O can hydrogenate sodium bicarbonate to
sodium formate with a TON of 610 (Table 7.5, entry 2) (164).
In addition to this, the same catalyst can be used for the hydrogenation
of carbon dioxide to methyl formate (TON 585), dimethylformamide
(TON 727), or formylpiperidine (TON 373). With the assistance of
Table 7.5 Nonprecious metals for the hydrogenation of carbon dioxide
Entry Catalyst Solvent Base/additive Temp. ( C) Pressure CO2/H2 (bar) Time (h) TON () TOF (h1) Refs.
1 FeCl3/dcpe DMSO DBU 50 60/40 7.5 113 15.1 (163)
2 Fe(BF4)26H2O, PP3 MeOH NaHCO3 80 0/60 20 610 n.a (164)
3 Co(BF4)26H2O, PP3 MeOH NaHCO3 80 0/60 20 3877 n.a (165)
a
4 trans-[(tBu-PNP)Fe(H2) H2O NaOH 80 3.33/6.66 5 788 n.a (166)
(CO)]
5 NiCl2(dcpe) DMSO DBU 50 160/40 216 4400 20 (163)
6 Ni(dppe)2 Benzene NEt3 RT 25/25 20 7 0.4 (72)
7 TiCl4/Mg THF None RT 1/1 24 15 n.a. (167)
8 MoCl3/dcpe DMSO DBU 50 60/40 7.5 63 8 (163)
a
THF as additive.
All abbreviations can be found in “List of the Used Abbreviations.”
246 Arno Behr and Kristina Nowakowski

Figure 7.9 Mechanism of the carbon dioxide hydrogenation with Fe(BF4)26H2O, PP3 (164).

in situ high-pressure NMR spectroscopy and X-ray analysis, the mechanism


of the iron-catalyzed reaction could be verified for the first time (Figure 7.9).
The active and well-defined iron hydride complexes [FeH(PP3)]BF4
and [FeH(H2)(PP3)]BF4 are formed under these reaction conditions
(Table 7.5, entry 2). Under carbon dioxide pressure, the intermediate
[FeH(CO2)(PP3)] is formed. Subsequently, the insertion of carbon dioxide
into the M–H bond occurs, and after a further intermediate stage, formic
acid is eliminated.
Apart from investigations on the Fe(BF4)26H2O system, an analogous
cobalt precursor was developed and tested (165). As a result of this exami-
nation, sodium formate was obtained in 94% yield by the hydrogenation
of sodium bicarbonate with the cobalt complex Co(BF4)26H2O
and the ligand PP3. No activity was observed with various other
common ligands, for example, xantphos, triphenylphosphine, and 1,2-bis
(diphenylphosphino)ethane. By increasing temperature to 120  C and
catalyst loading of 3.49  106 mol, a remarkable TON of 3877 was
reached. A variation of cobalt precursors can be performed in this case,
because several cobalt complexes, such as Co(acac)2, Co(acac)3, and CoCl2
in combination with the phosphorous ligand PP3, reached similar results. As
with the previously presented iron catalyst, also, derivates of formic acid can
be synthesized.
A further iron-based catalyst is the iron–pincer complex trans-[(tBu-
PNP)Fe(H2)(CO)] developed by Milstein et al. (see Figure 7.10) (166).
Under low pressures (6–10 bar), the hydrogenation of carbon dioxide
and sodium bicarbonate occurs in aqueous solution at 80  C with a TON
of 788 (Table 7.5, entry 4). The investigations of the mechanism by
NMR showed that the iron hydride directly attacks the carbon dioxide
and then the formate compound is replaced by water.
CO2 Chemistry 247

Figure 7.10 Iron–pincer complex (166).

3. CONTINUOUS HYDROGENATION OF CARBON


DIOXIDE IN MINIPLANT SCALE
In the previous chapters, the optimization of catalyst systems in batch
operation mode was described. In contrast, our research work is focused on
the homogeneously catalyzed hydrogenation of carbon dioxide to formic
acid performed in a continuously operated miniplant (168). In general, min-
iplants are used in process development for the preparation of feasibility
studies of new processes. By employing the miniplant technique, important
information of the reaction can be attained, such as the long-term stability of
the catalyst, the accumulation of by-products, and the catalyst losses in the
product phase. Further applications are the examination of the recycling
concepts and process optimization. Additionally, miniplants are used in
industry for the production of small production volumes, sample quantities,
and fine chemicals (169–171).
When using homogeneous catalysis, the recycling of the homogeneous
transition metals without a loss of activity or the resynthesis of the non-
active catalyst is a possible pathway for an economical process. To accom-
plish the catalyst recycling, several reaction pathways are possible, for
example, the use of liquid–liquid multiphase systems, the reaction with
in situ, or the postextraction of the product or the catalyst or thermo-
morphic multicomponent solvent systems (172–176). The most important
and industrially used application of the liquid–liquid multiphase concept is
the hydroformylation of propene to butanals in the Ruhrchemie/Rhône–
Poulenc process. The aqueous phase contains the water-soluble catalyst,
while the butanals and the propene exist in the organic phase (177,178).

3.1. Preliminary batch experiments


Prior to transferring the hydrogenation of carbon dioxide into the miniplant
technique, the reaction has been investigated and optimized in laboratory
scale via batch experiments. Various homogeneous transition metals with
248 Arno Behr and Kristina Nowakowski

different oxidations levels, especially ruthenium-, rhodium-, and iridium-


based catalysts, were applied in the hydrogenation of carbon dioxide to
formic acid in aqueous solution. The strong base triethylamine serves as
additive, and further longer-chain amines were also deployed in the hydro-
genation of carbon dioxide, in order to find out in what way the basicity of
the amine influences the catalytic activity of the catalyst (see Table 7.6).
Ruthenium(III)tris(acetylacetonate) (Ru(acac)3, Figure 7.11) proved to be
a robust and highly active catalyst in the hydrogenation of carbon dioxide.
The formed product is the salt consisting of formic acid and triethylamine.
Performing the hydrogenation of carbon dioxide in the presence of
Ru(acac)3, a TON up to 73,000 is achieved after a reaction time of 3 h
at 25  C under 30 bar CO2/H2 pressure (1:1). These reaction conditions
were optimized for the application in the continuously operated miniplant.
In addition, the optimization of the reaction pressure was performed in the
miniplant, where a continuous gas dosing is possible. As shown in Table 7.6,
the TON of the hydrogenation of carbon dioxide decreases with increasing
alkyl chain length of the amine. The effect observed can be ascribed to the
lower solubility of carbon dioxide, the decreasing pH level, and the lower
solubility of the longer-chain amines in water.

Table 7.6 Tertiary amines in the hydrogenation of carbon dioxide


Base TON [] pH []
1 Triethylamine 73,000 12.7
2 Tripropylamine 21,365 11.4
3 Tributylamine 11,500 10.6
4 Trihexylamine 2250 7.2
5 Trioctylamine 75 7
VReactor ¼ 300 ml, Vl ¼ 75 ml, p(H2) ¼ p(CO2), p ¼ 30 bar, cat: Ru(acac)3, T ¼ 25  C, c(cat.) ¼ 0.5 m-
mol l1, t ¼ 3 h, solvent ¼ water, D ¼ 700 rpm.

Figure 7.11 Ruthenium(III)tris(acetylacetonate).


CO2 Chemistry 249

As can be seen from Figure 7.12, the yield of formic acid can be increased by
the variation of the base concentration. The maximal TON is achieved with a
base concentration of about 3 mol l1. Higher base concentrations result in a
slight decrease of the TON. The reduction of the yield can be explained by the
lower water concentration, which stabilizes the formed product.
In addition, a two-phase system based on water and triethylamine, which
can be separated easily after the reaction, is obtained above a certain concen-
tration of triethylamine (5 mol l1) (see for the principle Figure 7.13). Inves-
tigations of the catalyst distribution in the aqueous and organic phase by

55,000

50,000

45,000

40,000

35,000
TON [–]

30,000

25,000

20,000

15,000

10,000

5000

0
0.1 0.5 1.0 1.1 2.1 3.1 4.1 5.1
c(NEt3)[mol/l]
Figure 7.12 Formation of formic acid–triethylamine salt depending on the base con-
centration. VReactor ¼ 30 ml, Vl ¼ 7,5 ml, p(H2) ¼ p(CO2), p ¼ 30 bar, cat: Ru(acac)3,
T ¼ 25  C, c(cat.) ¼ 0.5 mmol l1, t ¼ 6 h, solvent ¼ water, D ¼ 700 rpm.

Figure 7.13 Principle of the liquid–liquid multiphase systems.


250 Arno Behr and Kristina Nowakowski

means of inductively coupled plasma analysis (ICP) show that the majority of
the catalyst (95 w%) is dissolved in the organic phase. A remarkable advan-
tage of this reaction control is that no additional solvents are required for the
formation of a secondary liquid phase. However, the phase ratio is
depending on the yield of product. Before starting the reaction in batch,
the (NEt3)/(water) mass ratio is 1.6; during the reaction, it decreases to
0.4. Performing the reaction in the miniplant scale, constant phase ratios
are absolutely required.
The existing azeotrope between water and formic acid can be avoided by
extraction. For the extraction of formic acid from aqueous solution, different
aliphatic, long-chain amines can be used (179,180). For example, tri-
octylamine is particularly suitable for the extraction due to the
nonmiscibility with water and excellent extractive properties towards formic
acid (181–183). For the hydrogenation of carbon dioxide in the presence of
triethylamine, an extraction of the product is not possible due to the forma-
tion of a formic acid–triethylamine salt, which is only soluble in a protic
solvent.

3.2. Miniplant for carbon dioxide hydrogenation


In addition to the investigations in laboratory scale, a continuous and adjust-
able miniplant with three unit operations, reaction, liquid–liquid separation,
and liquid–gas separation, has been developed and built (Figure 7.14).
The reaction proceeds in a continuously stirred tank reactor, where
higher gas solubility is realized via a newly developed gas-entry stirrer.
Dosing the gaseous and liquid educts is performed by a gas sparger and
two piston pumps for water and triethylamine. After adjusting the

Figure 7.14 General flow sheet of the miniplant.


CO2 Chemistry 251

residence time in the reactor, the two phases will be separated in a liquid–
liquid separator. The organic amine phase containing the ruthenium cat-
alyst is recycled by means of a gear pump. These two unit operations
are held on the reaction pressure to enable the recirculation of the catalyst
phase. The miniplant is equipped with several pressure and temperature
sensors and is regulated via a fully automatic control system. Hereby, a high
operational safety is achieved. The reaction progress and the formation of
gaseous by-products can be traced via online gas chromatography.

3.3. Miniplant experiments


First, a successful scale-up of the hydrogenation of carbon dioxide was exe-
cuted into the miniplant technique. Similar results were achieved in the
miniplant with the laboratory reaction conditions. An increased yield can
be achieved by continuous gas supply as well as higher initial pressure
(see Figure 7.15).
The semicontinuous reaction control at 30 bar has been carried out;
however, a white solid precipitated due to the high product formation.
Based on this result, no TON could be calculated. In consideration of

80,000

70,000

60,000

50,000
TON [–]

40,000

30,000

20,000

10,000

0
10 20 30
Initialpressure [bar]
Semi-continuous Batch

Figure 7.15 Comparison of different initial pressures and reaction operations in


the hydrogenation of carbon dioxide in miniplant scale. VReactor ¼ 2 l, Vl ¼ 500 ml,
p(H2) ¼ p(CO2), cat: Ru(acac)3, T ¼ 25  C, c(cat.) ¼ 0.5 mmol l1, c(NEt3) ¼ 5 mol l1,
t ¼ 3 h, solvent ¼ water, D ¼ 700 rpm.
252 Arno Behr and Kristina Nowakowski

50,000

45,000

40,000

35,000

30,000
TON [–]

25,000

20,000

15,000

10,000

5000

0
Run 1 Run 2 Run 3 Run 4 Run 5

Figure 7.16 Recycling of the catalyst phase in the hydrogenation of carbon dioxide in
miniplant scale via batch experiments. VReactor ¼ 2 l, Vl ¼ 500 ml, p(H2) ¼ p(CO2), p ¼ 10
bar (semi-continuous), cat: Ru(acac)3, T ¼ 25  C, c(cat.) ¼ 0.5 mmol l1, c(NEt3) ¼
5 mol l1, t ¼ 3 h, solvent ¼ water, D ¼ 700 rpm.

the changing phase ratio depending on the product yield, the continuous
hydrogenation of carbon dioxide is carried out at 10 bar.
For a continuously operated process, constant activity of the catalyst is
desirable. Therefore, the recycling of the catalyst was examined. As can
be seen in Figure 7.16, the catalyst phase could be recycled in five consec-
utive runs in batch experiments with continuous gas dosing, and the activity
of the ruthenium catalyst decreases only slightly.
With regard to the recycling experiments, it was found that the catalyst
leaching depends on the yield of product. The more the product is formed,
the lower is the catalyst leaching into the aqueous product phase.
The miniplant concept developed could be verified by the first contin-
uous operations. In further investigations, the long-term stability of the cat-
alyst, the catalyst losses, and the accumulation of by-products will be
examined in more detail.

4. CONCLUSIONS
In the present chapter, a comprehensive overview of the homogenous
hydrogenation of carbon dioxide to formic acid and its derivatives has been
given. Various noble metal complexes based on rhodium, ruthenium, and
CO2 Chemistry 253

Figure 7.17 Formic acid–carbon dioxide cycle.

iridium were summarized and compared as catalyst. In recent years, the cat-
alytic activity of these homogeneous catalysts has been increased successfully.
In addition, intensive research in the hydrogenation of carbon dioxide with
nonprecious metals has been carried out. The iron and cobalt catalysts devel-
oped show in part similar results compared with the precious metal-based
catalysts. Latest research results present that the highest catalytic activity is
achieved by the Ir(III)–pincer complex (Figure 7.8C). In most cases, formic
acid derivates or formic acid salts are synthesized. Although several separa-
tion methods to obtain pure formic acid are pointed out, the separation of
the resulting products still remains a great challenge. The catalysts used for
the conversion of carbon dioxide also catalyze the decomposition of the
formed product. This result is on the one hand negative considering the sep-
aration; on the other hand, it is highly favorable in the terms of the hydrogen
storage in the carbon dioxide–formic acid cycle (see Figure 7.17). As men-
tioned in the introduction, formic acid can store - relating to the mass - less
hydrogen, but it is definitely a feasible possibility to store hydrogen. However,
the hydrogenation of carbon dioxide is only economical, if the required
hydrogen is produced by regenerative processes on the basis of water. The
presented concept is the research topic in the German joint research project
CO2RRECT (“CO2-Reaction using Regenerative Energies and Catalytic
Technologies”), in which the main focus of investigation is on the catalytic
conversion of carbon dioxide using regenerative energies to form formic
acid and the subsequent decomposition to hydrogen as well as carbon mon-
oxide. By this means, the produced formic acid is a compound, which can
easily be stored and transported for important basic chemicals.

ACKNOWLEDGMENTS
This research work is part of the German project “CO2RRECT,” in which 18 research
groups work together for an efficient use of carbon dioxide as a carbon building block.
The project is sponsored by the German Federal Ministry of Education and Research. In
addition, we thank the Umicore AG & Co. KG for catalyst donation.
254 Arno Behr and Kristina Nowakowski

REFERENCES
1. Behr, A. Chem. Ing. Tech. 1985, 57, 893.
2. Behr, A. Carbon Dioxide Activation by Metal Complexes. Wiley VCH: Weinheim, 1988.
3. Behr, A. Angew. Chem. 1988, 100, 681.
4. Cokoja, M.; Bruckmeier, C.; Rieger, B.; Herrmann, W. A.; Kühn, F. E. Angew. Chem.
2011, 123, 8662.
5. Omae, I. Catal. Today 2006, 115, 33.
6. Leitner, W. Angew. Chem. 1995, 107, 2391.
7. Aresta, M.; Dibenedetto, A. Dalton Trans. 2007, 28, 2975.
8. Federsel, C.; Jackstell, R.; Beller, M. Angew. Chem. 2010, 122, 6392.
9. Sakakura, T.; Choi, J. C.; Yasuda, H. Chem. Rev 2007, 107, 2365.
10. Aresta, M. Carbon Dioxide as Chemical Feedstock. Wiley-VCH: Weinheim, 2010.
11. Arakawa, H.; Aresta, M.; Armor, J. N.; Barteau, M. A.; Beckman, E. J.; Bell, A. T.;
Bercaw, E.; Creutz, C.; Dinjus, E.; Dixon, D. A.; Domen, K.; DuBois, D. L.;
Eckert, J.; Fujita, E.; Gibson, D. H.; Goddard, W. A.; Goodman, D. W.; Keller, J.;
Kubas, G. J.; Kung, H. H.; Lyons, J. E.; Manzer, L. E.; Marks, T. J.;
Morokuma, K.; Nicholas, K. M.; Periana, R.; Que, L.; Rostrup-Nielson, J.;
Sachtler, W. M. H.; Schmidt, L. D.; Sen, A.; Somorjai, G. A.; Stair, P. C.;
Stults, B. R.; Tumas, W. Chem. Rev. 2001, 101, 953.
12. Song, C. Catal. Today 2006, 115, 2.
13. Peters, M.; Köhler, B.; Kuckshinrichs, W.; Leitner, W.; Markewitz, P.; Müller, T. E.
ChemSusChem 2011, 4, 1216.
14. Sullivan, B. P.; Krist, K.; Guard, H. E. Electrochemical and Electrocatalytic Reactions of Car-
bon Dioxide. Elsevier Science: Amsterdam, 1993.
15. Willner, I.; Willner, B. Top. Curr. Chem. 1991, 153.
16. Hawecker, J.; Lehn, J.-M.; Ziessel, R. Helv. Chim. Acta 1986, 69, 1990.
17. Li, W. In Electrocatalytic Reduction of CO2 to Small Organic Molecule Fuels on Metal Cat-
alysts; Hu, Y. H. Ed.; Advances in CO2 Conversion and Utilization; Vol. 1056,
American Chemical Society: Washington, DC, 2010.
18. Ali, M. M.; Sato, H.; Mizukawa, T.; Tsuge, K.; Haga, M.; Tanaka, K. Chem. Commun.
1998, 249.
19. Ogata, T.; Yanagida, S.; Brunschwig, B. S.; Fujita, E. J. Am. Chem. Soc. 1995, 117,
6708.
20. Sato, S.; Morikawa, T.; Saeki, S.; Kajino, T.; Motohiro, T. Angew. Chem. 2010, 122,
5227.
21. Ayers, W. M. Ed.; Catalytic Activation of Carbon Dioxide; ACS Symposium Series; Vol.
363, American Chemical Society: Washington, DC, 1988.
22. Benson, E. E.; Kubiak, C. P.; Sathrum, A. J.; Smieja, J. M. Chem. Soc. Rev. 2009, 38,
89.
23. Drees, M.; Cokoja, M.; Kühn, F. E. ChemCatChem 2012, 4, 1703.
24. Bhargava, B. L.; Yasaka, Y.; Klein, M. L. J. Phys. Chem. B 2011, 115, 14136.
25. Behr, A.; Neuberg, S. Erdöl Erdgas Kohle 2009, 125, 367.
26. U.S. Energy Information Administration. www.eia.doe.gov (accessed 2013).
27. Arpe, H. J. Industrial Organic Chemistry, 5th ed.; Wiley-VCH: Weinheim, 2010.
28. Reutemann, W.; Kieczka, H. In Formic Acid. Ullmann’s Encyclopedia of Industrial Chem-
istry, Wiley-VCH: Weinheim, 2005.
29. Boddien, A.; Junge, H.; Beller, M. Nachrichten aus der Chemie 2011, 59, 1142.
30. Boddien, A.; Mellmann, D.; Gärtner, F.; Jackstell, R.; Junge, H.; Dyson, P. J.;
Laurenczy, G.; Ludwig, R.; Beller, M. Science 2011, 333, 1733.
31. Morris, D. J.; Clarkson, G. J.; Wills, M. Organometallics 2009, 28, 4133.
32. Fukuzumi, S.; Kobayashi, T.; Suenobu, T. J. Am. Chem. Soc. 2010, 132, 1496.
CO2 Chemistry 255

33. Himeda, Y. Green Chem. 2009, 11, 2018.


34. Himeda, Y.; Miyazawa, S.; Hirose, T. ChemSusChem 2011, 4, 487.
35. Fellay, C.; Dyson, P. J.; Laurenczy, G. Angew. Chem. 2008, 120, 4030.
36. Fellay, C.; Yan, N.; Dyson, P. J.; Laurenczy, G. Chem. Eur. J. 2009, 15, 3752.
37. Papp, G.; Csorba, J.; Laurenczy, G.; Joó, F. Angew. Chem. 2011, 123, 10617.
38. Joó, F. ChemSusChem 2008, 1, 805.
39. Boddien, A.; Loges, B.; Gärtner, F.; Torborg, C.; Fumino, K.; Junge, H.; Ludwig, R.;
Beller, M. J. Am. Chem. Soc. 2010, 132, 8924.
40. Scholten, J. D.; Prechtl, M. H. G.; Dupont, J. ChemCatChem 2010, 2, 1265.
41. Zhang, Y.; Zhang, J.; Zhao, L.; Sheng, C. Energy Fuels 2009, 24, 95.
42. Enthaler, S. ChemSusChem 2008, 1, 801.
43. Enthaler, S.; Langermann, J.; Schmidt, T. Energy Environ. Sci. 2010, 3, 1207.
44. Loges, B.; Boddien, A.; Gärtner, F.; Junge, H.; Beller, M. Top. Catal. 2010, 53, 902.
45. Boddien, A.; Loges, B.; Junge, H.; Gärtner, F.; Noyes, J. R.; Beller, M. Adv. Synth.
Catal. 2009, 351, 2517.
46. Jiang, Z.; Xiao, T.; Kuznetsov, V. L.; Edwards, P. P. Philos. Trans. R. Soc. A Math.
Phys. Eng. Sci. 2010, 368, 3343.
47. Denise, B.; Sneeden, R. P. A. ChemTech 1982, 108.
48. Tominaga, K.; Sasaki, Y.; Kawai, M.; Watanabe, T.; Saito, M. J. Chem. Soc. Chem.
Commun. 1993, 629.
49. Dorner, R. W.; Hardy, D. R.; Williams, F. W.; Willauer, H. D. Energy Environ. Sci.
2010, 3, 884.
50. Gomes, C. D. N.; Jacquet, O.; Villiers, C.; Thuery, P.; Ephritikhine, M.; Cantat, T.
Angew. Chem. Int. Ed. 2012, 51, 187.
51. Schmid, L.; Canonica, A.; Baiker, A. Appl. Catal. A 2003, 255, 23.
52. Wesselbaum, S.; vom Stein, T.; Klankermayer, J.; Leitner, W. Angew. Chem. 2012,
124, 7617.
53. Liao, F.; Huang, Y.; Ge, J.; Zheng, W.; Tedsree, K.; Collier, P.; Hong, X.; Tsang, S. C.
Angew. Chem. Int. Ed. 2011, 50, 2162.
54. Tominaga, K.; Sasaki, Y.; Watanabe, T.; Saito, M. Bull. Chem. Soc. Jpn. 1995, 68, 2837.
55. Tominaga, K.; Sasaki, Y.; Hagihara, K.; Watanabe, T.; Saito, M. Chem. Lett. 1994, 23,
1391.
56. Fiedler, E.; Grossmann, G.; Kersebohm, D. B.; Weiss, G.; Witte, C. In Methanol.
Ullmann’s Encyclopedia of Industrial Chemistry, Wiley-VCH: Weinheim, 2008.
57. Darensbourg, D. J.; Ovalles, C. J. Am. Chem. Soc. 1984, 106, 3750.
58. Lau, C. P.; Chen, Y. Z. J. Mol. Catal. A Chem. 1995, 101, 33.
59. Jessop, P. G.; Hsiao, Y.; Ikariya, T.; Noyori, R. J. Chem. Soc. Chem. Commun. 1995,
707.
60. Jessop, P. G.; Hsiao, Y.; Ikariya, T.; Noyori, R. J. Am. Chem. Soc. 1994, 116, 8851.
61. Kayaki, Y.; Suzuki, T.; Ikariya, T. Chem. Lett. 2001, 30, 1016.
62. Haynes, P.; Slaugh, L. H.; Kohnle, J. F. Tetrahedron Lett. 1970, 11, 365.
63. Inoue, Y.; Sasaki, Y.; Hashimoto, H. J. Chem. Soc. Chem. Commun. 1975, 718.
64. Jessop, P. G. In Homogeneous Hydrogenation of Carbon Dioxide; de Vries, J. G.;
Elsevier, C. J.; de Vries, J. G.; Elsevier, C. J. Eds.; Handbook of Homogeneous Hydro-
genation; Vol. 1, Wiley-VCH: Weinheim, 2007.
65. Tsai, J. C.; Nicholas, K. M. J. Am. Chem. Soc. 1992, 114, 5117.
66. Jessop, P. G.; Morris, R. H. Coord. Chem. Rev. 1992, 121, 155.
67. Musashi, Y.; Sakaki, S. J. Am. Chem. Soc. 2000, 122, 3867.
68. Urakawa, A.; Iannuzzi, M.; Hutter, J.; Baiker, A. Chem. Eur. J. 2007, 13, 6828.
69. Bo, C.; Dedieu, A. Inorg. Chem. 1989, 28, 304.
70. Jessop, P. G.; Ikariya, T.; Noyori, R. Chem. Rev. 1995, 95, 259.
71. Karsch, H. H. Chem. Ber. 1977, 110, 2213.
256 Arno Behr and Kristina Nowakowski

72. Inoue, Y.; Izumida, H.; Sasaki, Y.; Hashimoto, H. Chem. Lett. 1976, 5, 863.
73. Yin, C.; Xu, Z.; Yang, S. Y.; Ng, S. M.; Wong, K. Y.; Lin, Z.; Lau, C. P. Organome-
tallics 2001, 20, 1216.
74. Leitner, W.; Dinjus, E.; Gaßner, F. J. Organomet. Chem. 1994, 475, 257.
75. Dwars, T.; Oehme, G. Adv. Synth. Catal. 2002, 344, 239.
76. Pinault, N.; Bruce, D. W. Coord. Chem. Rev. 2003, 241, 1.
77. Herrmann, W. A.; Kohlpaintner, C. W. Angew. Chem. Int. Ed. 1993, 32, 1524.
78. Laurenczy, G.; Joó, F.; Nádasdi, L. Inorg. Chem. 2000, 39, 5083.
79. Elek, J.; Nádashi, L.; Papp, G.; Laurenczy, G.; Joó, F. Appl. Catal. A 2003, 255, 59.
80. Joó, F.; Nádasdi, L.; Elek, J.; Laurenczy, G. Chem. Commun. 1999, 971.
81. Joó, F.; Laurenczy, G.; Karady, P.; Elek, J.; Nádasdi, L.; Roulet, R. Appl. Organomet.
Chem. 2000, 14, 857.
82. Farlow, M. W.; Adkins, H. J. Am. Chem. Soc. 1935, 57, 2222.
83. Baiker, A. Appl. Organomet. Chem. 2000, 14, 751.
84. Zhang, Y.; Fei, J.; Yu, Y.; Zheng, X. Catal. Commun. 2004, 5, 643.
85. Yu, Y. M.; Fei, J. H.; Zhang, Y. P. Chin. Chem. Lett. 2006, 261.
86. Jessop, P. G.; Ikariya, T.; Noyori, R. Nature 1994, 368, 231.
87. Jessop, P. G.; Hsiao, Y.; Ikariya, T.; Noyori, R. J. Am. Chem. Soc. 1996, 118, 344.
88. Hao, C.; Wang, S.; Li, M.; Kang, L.; Ma, X. Catal. Today 2011, 160, 184.
89. Fachinetti, G.; Preti, D. Process for Preparing Formic Acid. U.S. 20,130,006,015,
2013.
90. Preti, D.; Resta, C.; Squarcialupi, S.; Fachinetti, G. Angew. Chem. 2011, 50, 12551.
91. Preti, D.; Squarcialupi, S.; Fachinetti, G. ChemCatChem 2012, 4, 469.
92. Himeda, Y.; Onozawa-Komatsuzaki, N.; Sugihara, H.; Arakawa, H.; Kasuga, K. In
50th Symposium on Organometallic Chemistry, 2003.
93. Munshi, P.; Main, A. D.; Linehan, J. C.; Tai, C. C.; Jessop, P. G. J. Am. Chem. Soc.
2002, 124, 7963.
94. Gao, Y.; Kuncheria, J. K.; Jenkins, H. A.; Puddephatt, R. J.; Yap, G. P. A. Dalton Trans.
2000, 3212.
95. Zhang, J. Z.; Li, Z.; Wang, H.; Wang, C. Y. J. Mol. Catal. A Chem. 1996, 112, 9.
96. Man, M. L.; Zhou, Z.; Ng, S. M.; Lau, C. P. Dalton Trans. 2003, 3727.
97. Ng, S. M.; Yin, C.; Yeung, C. H.; Chan, T. C.; Lau, C. P. Eur. J. Inorg. Chem. 2004,
2004, 1788.
98. Yamaji, T. Japan Kokai Tokkyo Koho 1981, 140948.
99. Chu, H. S.; Lau, C. P.; Wong, K. Y.; Wong, W. T. Organometallics 1998, 17, 2768.
100. Drury, D.J.; Hamlin, J.E. Production of Formate Salts. EP0095321, 1983.
101. Behr, A.; Ebbinghaus, P.; Naendrup, F. Chem. Ing. Tech. 2003, 75, 877.
102. Taqui Khan, M. M.; Halligudi, S. B.; Shukla, S. J. Mol. Catal. 1989, 57, 47.
103. Sanz, S.; Azua, A.; Peris, E. Dalton Trans. 2010, 39, 6339.
104. Himeda, Y.; Onozawa-Komatsuzaki, N.; Sugihara, H.; Arakawa, H.; Kasuga, K.
Organometallics 2004, 23, 1480.
105. Federsel, C.; Jackstell, R.; Boddien, A.; Laurenczy, G.; Beller, M. ChemSusChem 2010,
3, 1048.
106. Himeda, Y.; Onozawa-Komatsuzaki, N.; Sugihara, H.; Kasuga, K. Organometallics
2007, 26, 702.
107. Thai, T. T.; Therrien, B.; Süss-Fink, G. J. Organomet. Chem. 2009, 694, 3973.
108. Challand, N.; Sava, X.; Röper, M. Verfahren zur Herstellung von Ameisensäure.
WO2008/116799, 2008.
109. Hayashi, H.; Ogo, S.; Abura, T.; Fukuzumi, S. J. Am. Chem. Soc. 2003, 125, 14266.
110. Hayashi, H.; Ogo, S.; Fukuzumi, S. Chem. Commun. 2004, 2714.
111. Ogo, S.; Kabe, R.; Hayashi, H.; Harada, R.; Fukuzumi, S. Dalton Trans. 2006, 4657.
112. Laurenczy, G.; Joó, F.; Nádasdi, L. Int. J. High Pres. Res. 2000, 18, 251.
CO2 Chemistry 257

113. Ohnishi, Y.; Matsunaga, T.; Nakao, Y.; Sato, H.; Sakaki, S. J. Am. Chem. Soc. 2005,
127, 4021.
114. Jessop, P. G.; Leitner, W. Chemical Synthesis Using Supercritical Fluids. Wiley-VCH:
Weinheim, 2008.
115. Anastas, P. T. Ed.; Handbook of Green Chemistry; Leitner, W.; Jessop, P. G. Eds.; Green
Solvent, Vol. 4: Supercritical Solvents Wiley-VCH: Weinheim, 2010.
116. Urakawa, A.; Jutz, F.; Laurenczy, G.; Baiker, A. Chem. Eur. J. 2007, 13, 3886.
117. Tsang, C. Y.; Street, W. B. Chem. Eng. Sci. 1981, 36, 993.
118. Getty, A. D.; Tai, C. C.; Linehan, J. C.; Jessop, P. G.; Olmstead, M. M.;
Rheingold, A. L. Organometallics 2009, 28, 5466.
119. Wesselbaum, S.; Hintermair, U.; Leitner, W. Angew. Chem. 2012, 51, 8585.
120. Leitner, W.; Hintermair, U.; Wesselbaum, S. CO2 Hydrogenation Method for Pro-
ducing Formic Acid, WO 2012/095345, 2012.
121. Zhang, Z.; Hu, S.; Song, J.; Li, W.; Yang, G.; Han, B. ChemSusChem 2009, 2, 234.
122. Yasaka, Y.; Wakai, C.; Matubayasi, N.; Nakahara, M. J. Phys. Chem. A 2010, 114,
3510.
123. Yamaji, T. Japan Kokai Tokkyo Koho 1981, 166.146.
124. Ezhova, N. N.; Kolesnichenko, N. V.; Bulygin, A. V.; Slivinskii, E. V.; Han, S. Russ.
Chem. Bull. 2002, 51, 2165.
125. Graf, E.; Leitner, W. Chem. Commun. 1992, 8, 623.
126. Burgemeister, T.; Kastner, F.; Leitner, W. Angew. Chem. 1993, 105, 781.
127. Dietrich, J.; Schindler, S. Z. Anorg. Allg. Chem. 2008, 634, 2487.
128. Hutschka, F.; Dedieu, A.; Leitner, W. Angew. Chem. Int. Ed. 1995, 34, 1742.
129. Hutschka, F.; Dedieu, A.; Eichberger, M.; Fornika, R.; Leitner, W. J. Am. Chem. Soc.
1997, 119, 4432.
130. Pomelli, C. S.; Tomasi, J.; Solà, M. Organometallics 1998, 17, 3164.
131. Gassner, F.; Leitner, W. Chem. Commun. 1993, 19, 1465.
132. Fornika, R.; Görls, H.; Seemann, B.; Leitner, W. J. Chem. Soc. Chem. Commun. 1995,
1479.
133. Leitner, W.; Dinjus, E.; Gaßner, F. In Aqueous-Phase Organometallic Catalysis—
Concepts and Applications; Cornils, B.; Herrmann, W. A. Eds.; Wiley VCH:
Weinheim, 1998.
134. Lindner, E.; Keppeler, B.; Wegner, P. Inorg. Chim. Acta 1997, 258, 97.
135. Karakhanov, E. A.; Egarzar’yants, S. V.; Kardashev, S. V.; Maksimov, A. L.;
Minos’yants, S. S.; Sedykh, A. D. Petrol. Chem. 2001, 41, 268.
136. Zou, F.; Cole, J. M.; Jones, T. G. J.; Jiang, L. Appl. Organomet. Chem. 2012, 546.
137. Zhao, G.; Joó, F. Catal. Commun. 2011, 14, 74.
138. Jószai, I.; Joó, F. J. Mol. Catal. A Chem. 2004, 224, 87.
139. Musashi, Y.; Sakaki, S. J. Am. Chem. Soc. 2002, 124, 7588.
140. Preti, D.; Squarcialupi, S.; Fachinetti, G. Angew. Chem. 2010, 120, 2635.
141. Barrow, G. M.; Yerger, E. A. J. Am. Chem. Soc. 1954, 76, 5211.
142. Wagner, K. Angew. Chem. Int. Ed. 1970, 9, 50.
143. Gladiali, S.; Alberico, E. Chem. Soc. Rev. 2005, 35, 226.
144. Sakamoto, M.; Shimizu, I.; Yamamoto, A. Organometallics 1994, 13, 407.
145. Anderson, J.J.; Hamlin, J.E. Verfahren zur Herstellung von Ameisensäure. EP0126524,
1984.
146. Anderson, J.J.; Drury, D.J.; Hamlin, J.E. EP0181078, 1985.
147. Schaub, T.; Paciello, R. A. Angew. Chem. 2011, 123, 7416.
148. Schaub, T.; Paciello, R.; Mohl, K.-D.; Schneider, D.; Schäfer M.; Rittinger, S. Ver-
fahren zur Herstellung von Ameisensäure. EP2010/058208, 2010.
149. Hohenschutz, H.; Kiefer, H.; Schmidt J. Isolation of Formic Acid from its Aqueous,
Solutions. U.S. 4,217,460, 1980.
258 Arno Behr and Kristina Nowakowski

150. Kiefer, H.; Hupfer, L.; Lippert, F. Preparation of Formic Acid by Thermal Cleavage of
Quaternary Ammonium Formates. U.S. 5,294,740, 1993.
151. Kudo, K.; Sugita, N.; Takezaki, Y. Nihon Kagaku Kaishi 1977, 302.
152. Hashimoto, H.; Inoue, Y. Kokai Tokkyo Koho 1978, 7612.
153. Himeda, Y.; Onozawa-Komatsuzaki, N.; Sugihara, H.; Kasuga, K. J. Am. Chem. Soc.
2005, 127, 13118.
154. Himeda, Y. Eur. J. Inorg. Chem. 2007, 2007, 3927.
155. Himeda, Y. In Utilization of Carbon Dioxide as a Hydrogen Storage Material: Hydrogenation
of Carbon Dioxide and Decomposition of Formic Acid Using Iridium Complex Catalysts;
Hu, Y. H. Ed.; ACS Symposium Series; Vol. 1056, Oxford University Press:
Washington, DC, 2010.
156. Himeda, Y. Prep. Pap. Am. Chem. Soc. Div. Fuel Chem. 2012, 57, 246.
157. Moret, S.; Laurenczy, G.; Dyson, P. J. Dalton Trans. 2013, 42, 4353.
158. Azua, A.; Sanz, S.; Peris, E. Chem. Eur. J. 2011, 17, 3963.
159. Tanaka, R.; Yamashita, M.; Nozaki, K. J. Am. Chem. Soc. 2009, 131, 14168.
160. Ahlquist, M. S. G. J. Mol. Catal. A Chem. 2010, 324, 3.
161. Yang, X. ACS Catal. 2011, 1, 849.
162. Evans, G. O.; Newell, C. J. Inorg. Chim. Acta 1978, 31, L387.
163. Tai, C. C.; Chang, T.; Roller, B.; Jessop, P. G. Inorg. Chem. 2003, 42, 7340.
164. Federsel, C.; Boddien, A.; Jackstell, R.; Jennerjahn, R.; Dyson, P. J.; Scopelliti, R.;
Laurenczy, G.; Beller, M. Angew. Chem. 2010, 122, 9971.
165. Federsel, C.; Ziebart, C.; Jackstell, R.; Baumann, W.; Beller, M. Chem. Eur. J. 2012,
18, 72.
166. Langer, R.; Diskin-Posner, Y.; Leitus, G.; Shimon, L. J. W.; Ben-David, Y.;
Milstein, D. Angew. Chem. 2011, 50, 9948.
167. Jezowska-Trzebiatowska, B.; Sobota, P. J. Organomet. Chem. 1974, 80, C27.
168. Nowakowski, K. TU Dortmund. PhD Thesis, 2014.
169. Behr, A.; Ebbers, W.; Wiese, N. Chem. Ing. Tech. 2000, 72, 1157.
170. Behr, A.; Witte, H.; Zagajewski, M. Chem. Ing. Tech. 2012, 84, 694.
171. Deibele, L.; Dohrn, R. Grundsätze der Miniplant-Technik. Wiley-VCH: Weinheim,
2006.
172. Behr, A.; Henze, G.; Johnen, L.; Awungacha, C. J. Mol. Catal. A Chem. 2008, 285, 20.
173. Behr, A.; Obst, D.; Turkowski, B. J. Mol. Catal. A Chem. 2005, 226, 215.
174. Behr, A.; Roll, R. J. Mol. Catal. A Chem. 2005, 239, 180.
175. Behr, A.; Fängewisch, C. J. Mol. Catal. A Chem. 2003, 197, 115.
176. Behr, A.; Roll, R. Chem. Ing. Tech. 2005, 77, 748.
177. Behr, A. Angewandte homogene Katalyse. Wiley-VCH: Weinheim, 2008.
178. Behr, A.; Neubert, P. Applied Homogeneous Catalysis. Wiley-VCH: Weinheim, 2012.
179. Sahin, S.; Bayazit, S. S.; Bilgin, M.; Inci, I. J. Chem. Eng. Data 2010, 55, 1519.
180. Wardell, J. M.; King, C. J. J. Chem. Eng. Data 1978, 23, 144.
181. Hong, Y. K.; Hong, W. H.; Chang, Y. K. Biotechnol. Bioprocess Eng. 2001, 6, 347.
182. Li, Z.; Qin, W.; Dai, Y. J. Chem. Eng. Data 2003, 48, 1113.
183. Qin, W.; Li, Z.; Dai, Y. Ind. Eng. Chem. Res. 2003, 42, 6196.

You might also like