You are on page 1of 18

Review

Synthesis of Acrylic Acid Derivatives


from CO2 and Ethylene
Xiao Wang,1,2,* Hui Wang,1,* and Yuhan Sun1,3,*

The synthesis of acrylic acid and its derivatives with the use of CO2 and ethylene The Bigger Picture
feedstock is an exciting yet extremely challenging reaction. In addition to Acrylic and its derivatives are
the inherent stability of CO2, other challenges arise, such as the elimination important raw materials for the
of b-hydride from metallalactone to release free acrylate. Since the ground- manufacture of a wide variety of
breaking work more than three decades ago, tremendous efforts have been industrial and consumer products.
made to realize an effective catalytic system. From historical findings to recent In the theme of sustainability,
progress, this review gives a comprehensive overview of this reaction. Reac- more economical methods that
tions mediated by different transition-metal complexes and mechanistic details utilize cheaper and less toxic
are discussed in depth. feedstock are urgently needed.
Clean pathways can be realized
INTRODUCTION through the direct
Acrylic acid and its derivatives are extremely important building blocks for functionalization of CO2, which
the manufacture of various industrial and consumer products, such as plastics, plays a dual role of reducing the
textiles, coatings, elastomers, adhesives, and paints.1,2 The global annual market CO2 emission rate and producing
demand for sodium acrylate alone is 4 million tons. The world capacity of chemicals with added value.
acrylic acid is 6 million tons as of 2013 and is still rapidly increasing. Traditionally, Because ethylene is one of the
acrylates have been synthesized exclusively via a two-step oxidation of propene most abundant industrial
with O2 (via a transient acrolein intermediate) over heterogeneous catalysts feedstocks, manufacturing acrylic
at high temperature, in almost quantitative yield.3–5 As an alternative to oxida- acid and acrylates directly from
tive processes, carboxylation of unsaturated hydrocarbon using C1 feedstock CO2 and ethylene becomes both
has also garnered much attention. The most popular processes utilize CO and scientifically fascinating and
H2O as the C1 source with nickel6,7 or palladium8,9 complexes as catalysts; industrially relevant. As a dream
some processes, such as the Reppe process, have been shown to be commer- reaction in catalysis and CO2
cially viable.10,11 The acid-catalyzed Koch reaction is also an important process utilization, this challenging
to achieve direct carboxylation of alkenes.12 However, for olefinic substrates, transformation has drawn close
most of these methods produce saturated aliphatic acids rather than their un- attention for nearly 40 years. With
saturated counterparts. Moreover, although many of these processes are consid- continuous efforts from both
ered clean routes to manufacture carboxylic acids (including acrylic acid), more academia and industry, scientists
economical methods utilizing cheaper and less toxic feedstock are urgently are getting close to realizing an
needed. effective catalytic system. This
review covers critical discoveries
From a low carbon point of view, one fundamental strategy to reduce the CO2 and discusses possible future
emission rate is by means of carbon capture and utilization (CCU) through direct directions in this area.
functionalization of CO2, which serves a dual purpose of reducing greenhouse gas
emissions and producing value-added chemicals. Thus, ideally, it would be attrac-
tive to explore the preparation of acrylates with CO2 as an abundant, cheap, and sus-
tainable C1 feedstock. Manufacturing acrylates via the addition of ethylene and
CO2 is challenging and considered one of the ‘‘dream reactions’’ in catalysis
(Scheme 1). The thermodynamic stability of CO2 has led to challenges in its utiliza-
tion as a feedstock in synthetic applications. Few reactions and catalysts have
been developed to enable the straightforward conversion of CO2 to other chemi-
cals,13,14 and even fewer have been industrialized.15 Nevertheless, since pioneering

Chem 3, 211–228, August 10, 2017 ª 2017 Elsevier Inc. 211


catalyst O O
CO2 + or
reagent HO MO

Scheme 1. Synthesis of Acrylate from CO2 and Ethylene

work in the late 1970s, impressive progress has been made toward establishing an
effective CO2-to-acrylate process.16,17

This review covers research on the direct functionalization of alkenes (ethylene


in particular) with the use of CO2 as the carboxylation reagent. General principles
and key steps of this transformation are discussed in the context of the different
transition-metal catalysts used. These reactions are inherently important, and a
good understanding of them will inspire solutions to related chemical reactions uti-
lizing CO2 as building blocks.

Nickel-Mediated Reactions
Ligand-modified nickel(0) complexes are by far the most intensively investigated pro-
moters for the coupling of CO2 with ethylene. In 1982, Hoberg and Schaefer discov-
ered that in the presence of 2,20 -bipyridine (bpy), Ni(0) could undergo an addition
reaction with norbornene and CO2 to form a five-membered adduct (known as
Hoberg’s complex or nickelalactone) in 93% yield at 60 C (Scheme 2).18 The com-
plex could be hydrolyzed to give norbornane-2-carboxylic acid in 95% yield. In a
follow-up study, (dicyclohexylphosphino)ethane (dcpe) was used at the same tem-
perature (60 C), and saturated acid product was obtained in 80% yield.19 The addi-
tion reaction of CO2 with cyclopentene was also investigated by Hoberg et al.20

At almost the same time, Walther and Dinjus21 and Walther et al.22 reported
the Ni(bpy) and Ni(TMEDA)-mediated (TMEDA = N,N,N0 ,N’-tetramethylethylene-
diamine) addition of 2,3-dimethylbutadiene and CO2. The structure of the
7-membered nickelalactone formed was confirmed by X-ray analysis. When treated
with HX (X = halide) or H2SO4, the nickelalactone could be hydrolyzed to give
3,4-dimethylpent-3-enoic acid (Scheme 3). However, when treated with MeI, the ma-
jor product was 3,4-dimethylhex-3-enoic acid as a result of C alkylation. Additional O
alkylation was only observed when excessive MeI was used.

The first example of Ni-mediated CO2-ethylene addition was reported in 1983


by Hoberg and Schaefer, who used a stoichiometric amount of nickel complex
with bipyridine or (dicyclohexylphosphino)ethane (dcpe) as ligand.23 The Ni com-
plex underwent addition with CO2 and ethylene to form nickelalactone, which could
be solvolyzed under acidic conditions to afford propionic acid or its methyl ester 1CAS Key Lab of Low-Carbon Conversion Science
(Scheme 4). In 1987, a nickelalactone/1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) and Engineering, Shanghai Advanced Research
Institute, Chinese Academy of Sciences, 100
complex was isolated and structurally confirmed by X-ray analysis.24 However, these Haike Road, Pudong, Shanghai 201210, People’s
nickelalactones were not able to complete b-H elimination to release acrylic acid or Republic of China
its salt. 2Harvard NeuroDiscovery Center, Harvard
Medical School and Brigham & Women’s
Hospital, 65 Landsdowne Street, Cambridge, MA
The rationalization for the inconvertibility of nickelalactone to acrylate was pro- 02139, USA
vided by Graham et al.,25 who performed in-depth density functional theory (DFT) 3School of Physical Science and Technology,
computations by using density functional B3LYP with LANL2DZ (for nickel) and ShanghaiTech University, Shanghai 200031,
People’s Republic of China
6-31+G(d,p) (for other atoms) basis sets. It was argued that the catalytic reaction
*Correspondence: wangxiao@sari.ac.cn (X.W.),
with Hoberg-type nickelalactone ligated with bis-mDBU (a model DBU ligand) has wanghh@sari.ac.cn (H.W.),
three primary energy barriers: nickelacycle formation (+121.8 kJ/mol), b-H elimi- sunyh@sari.ac.cn (Y.S.)
nation (+147.4 kJ/mol), and reductive elimination to yield acrylic acid through a http://dx.doi.org/10.1016/j.chempr.2017.07.006

212 Chem 3, 211–228, August 10, 2017


Scheme 2. Hoberg’s CO2-Norbornene Addition Reaction

five-membered ring transition state (+104.1 kJ/mol) (Figure 1). The solvent-cor-
rected Gibbs free energy of the overall coupling process was strongly unfavorable
as well (+42.7 kJ/mol). Pápai et al.26 also performed DFT calculations for a series
of nickelalactone with substituted bpy ligands on the B3LYP/6-311++G(d,p) level
of theory. The computed activation barrier for the nickelalactone formation was
80–110 kJ/mol.

Nevertheless, Hoberg et al.27 demonstrated that b-hydride elimination could occur


smoothly in reactions with styrene. Mediated by the Ni(DBU) complex, styrene and
CO2 underwent an addition reaction to form a pair of regioisomeric nickalactones,
which could be hydrolyzed under aqueous acidic conditions to afford the saturated
acids (Scheme 5). The a-phenyl nickelalactone could undergo a thermo-b-hydride
elimination at 60 C, followed by hydrolysis or reductive elimination, to form cin-
namic acid in 3% isolated yield. At a higher temperature (85 C), the yield of cinnamic
acid drastically increased to 70%.

An important breakthrough was made in 2006, when Fischer et al.28 found that
b-hydride elimination of a nickelalactone could be achieved if the flexible and
compact bidentate ligand bis(dicyclohexylphosphino)methane (dppm) was used.
Other related studies by this group include Ni-mediated addition of CO2 with
dicyclopentadiene29 or alkyne,30 and TMEDA-pyridine ligand exchange for nickel-
alactone and ring-opening reactions by disulfides.31 In their research on the
Ni-dppm complex, the reactive intermediate [(dppm)2Ni(CH2CH2COO)] under-
went two distinct transformations: (1) b-hydride elimination to give a NiH(acrylate)
(dppm)2 complex and (2) reductive decoupling to give Ni(dppm)3, CO2, and
ethylene (Scheme 6). In the first pathway, however, the expected reductive elimi-
nation to release free acrylic acid was not successful. Instead, NiH(acrylate)
(dppm)2 combined with Ni(dppm)3 to form a binuclear Ni(I) complex with bridging
acrylate, dppm, and anionic diphenylphosphido groups, releasing one molecule of
Ph2PMe as confirmed by 31P nuclear magnetic resonance (NMR). The binuclear

Scheme 3. Walther’s CO2-Diene Addition Reaction

Chem 3, 211–228, August 10, 2017 213


Scheme 4. Hoberg’s CO2-Ethylene Addition Reaction and the Formation of Nickelalactone

complex was isolated, and the structure was confirmed by single-crystal X-ray anal-
ysis. Although the primary goal of producing acrylic acid was not achieved, this
study illustrated that the crucial step of b-hydride elimination could be realized
by choosing the right ligand.

Because geometric constraints cause b-hydride elimination to be the primary


obstacle to achieving the desired transformation, fragmentation of the five-
membered ring to generate an acyclic structure should facilitate elimination.
One approach is to add an exogenous electrophile to cause metallacycle
fragmentation and enable b-hydride elimination. After the observation of the
concomitant O-methylation with excessive MeI in the reaction of diene,21,22
in 2010 Bruckmeier et al.32 reported a method that utilized MeI to facilitate
the breaking of the Ni–O bond from the 5-membered nickelalactone. After ring
opening, the open-chain complex was more prone to b-hydride elimination to
yield methyl acrylate (Scheme 7). The transformation required 2–10 equiv of
MeI, with 10 equiv of MeI resulting in yields as low as 20%. With MeI as solvent,
the yield only slightly increased to 30%. Although not mentioned in the original
paper, we argue that the low yield was possibly caused by competing or
even more favorable C alkylation, which was reported by others.33–36 Moreover,
we hold a different view regarding the use of MeI, which is indeed not an ideal
methylating reagent because of its volatility and carcinogenicity, as well as the
introduction of the reductive and unremovable iodide. The need for excessive
MeI was later explained by a computational study,37 which suggested the poor
polarity of MeI rendered it unable to stabilize the key intermediates by forming
an ion pair.

Figure 1. Buntine’s Calculated Free Energy of Intermediates in a Ni(mDBU)2-Facilitated Reaction

214 Chem 3, 211–228, August 10, 2017


Scheme 5. Synthesis of Cinnamic Acid from CO2 and Styrene

Later, Lee et al.38 demonstrated that the diamine or diphosphine ligand of nickela-
lactones had a significant effect on the yield of methyl acrylate. Nickel ligated with
TMEDA or dppe gave the best yield of around 40%. To overcome the high reaction
barriers of b-H elimination, a large excess of methyl iodide is necessary. With the
same bidentate ligand, the yield only increased by 20% when the amount of MeI
was increased from 10 to 100 equiv. For a mixture of nickelalactone and 100 equiv
MeI, a kinetic study was performed by monitoring the shift in the n(C=O) bond
stretch with attenuated total reflectance-Fourier transform-infrared (ATR-FT-IR)
spectroscopy. A new IR band was observed at around n = 1,680 cm1, representing
complex B, which was in accord with the calculated value (1,720 G 30 cm1)
(Scheme 8). A slightly lower IR band of 1,672 cm1 was observed after 3 hr, indicating
the formation of p-bonded h2-methyl acrylate (C).

In contrast to the TMEDA complex that studied by Lee et al.,38 Plessow et al.39
proposed an alternative mechanism for MeI-assisted NiO bond cleavage of the
1,2-bis(di-tert-butylphosphino)ethane (dtbpe) ligated nickelalactone complex. This

Ph Ph
H
Ph2P P
Ni
P O Ph Ph
Ph2P O Ph2P PPh2
Ph Ph Ph Ph P
Ph2P PPh2
Ph2P P Ni Ni
Ni
P O O O O PPh2
Ph2P Ph2P
Ph Ph Ph Ph Ph
Ph PPh2
P P
bis(dppm)nickelalactone Ni (isolated)
P P PPh2
Ph Ph Ph Ph + Ph2PMe
(isolated) (detected)

Scheme 6. Formation of Walther’s Binuclear Ni(I) Complex

Chem 3, 211–228, August 10, 2017 215


Scheme 7. Hypothetical Catalytic Cycle of Rieger’s Methylation of Nickelalactone and Synthesis
of Methyl Acrylate

difference in mechanism again indicated that the ligand plays a crucial role in
determining the kinetic profile of the reaction. In a pathway similar to that pro-
posed by Lee et al., the transition state corresponding to the addition of MeI to
dtbpe-nickelalactone complex was too high in energy (DG = 271 kJ/mol) and was
considered less likely to be a major pathway for the methylation step (Scheme 9).
Instead, methylation via SN2 methylation of the carbonyl followed by ring opening
by the iodide should be the most probable pathway (DG = 73 kJ/mol). Methylation
on the oxygen of the NiO bond is not likely either because the activation energy
(DG = 111 kJ/mol) is much higher than the methylation of the carbonyl group.

To confirm the carbonyl-methylation pathway, single crystals of B1-BArF were ob-


tained by reacting complex A1 with MeOTf at 78 C and subsequent exchange
of counterions from TfO to BArF (Scheme 10).

A B

Scheme 8. Reaction Mechanism with the Ni-TMEDA Complex Proposed by Kühn

216 Chem 3, 211–228, August 10, 2017


Scheme 9. Proposed Reaction Pathways with the Ni-dtbpe Complex by Limbach and Hoffmann

In 2012, the Limbach group at Heidelberg University and researchers at BASF devel-
oped the first Ni-catalyzed acrylate formation reaction.40 Assisted by DFT-computed
DG values, they targeted the formation of sodium acrylate instead of free acrylic
acid, because the first energy barrier (nickelalactone formation) was reduced
(+101 versus +122 kJ/mol), and the overall process is exothermic for sodium acrylate
synthesis (DG = 59 kJ/mol). With dtbpe as the optimal ligand, the rate-determining
step was identified as base-facilitated deprotonation elimination at the a carbon of
the nickelalactone, rendering the p complex (Scheme 11). A strong base such as
t
BuONa was necessary to promote the a deprotonation. At 40 bar of CO2 and
5 bar of ethylene, a landmark turnover number (TON) of greater than 10 was
achieved. Limbach also discovered that the cation and the aggregation state of
the base play a crucial role in this reaction, as an appropriate cation could effectively
activate the carbonyl. Although the TON was nowhere near being applicable, it rep-
resents the first effective catalytic reaction, and a giant leap toward realizing an
industrially relevant process.

Because of the inherent incompatibility of strong bases (e.g., tBuONa) with CO2 or
carbonyl-activating reagents (e.g., Lewis acids),14 seeking less basic conditions to
adapt to the CO2 atmosphere was strongly desirable. In their follow-up research,
Huguet et al.41 replaced sodium alcoholates with weak bases such as substituted
sodium phenoxides, where sodium 2-fluorophenoxide was found to be the optimal
base. The dtbpe ligand in their previous catalytic system was replaced by the P-chi-
rogenic ligand BenzP*. An increased TON of 39 was achieved, and the pressure of
CO2 was reduced to as low as 10 bar, whereas in the Ni(dtbpe)-catalyzed reaction
it was three times higher (40 bar) (Scheme 12). Interestingly, it was found that a
greater amount of phenoxide and the addition of finely powdered zinc both further
increased the TON. A TON of 107 was realized with the addition of excessive zinc
(Zn/Ni = 100). Although the mechanism was not fully elucidated, it is suspected
that zinc reduces the Ni(II) intermediate species, thus facilitating the overall rate of
the reaction. Practically speaking, the necessity of excessive zinc is suitable for a

Chem 3, 211–228, August 10, 2017 217


Scheme 10. Obtaining the Crystal Structure for the Methylated Nickelalactone
Reprinted with permission from Lee et al. 38 Copyright 2013 American Chemical Society.

proof of concept but might not be industrially relevant and should preferably be
avoided.

Despite the unprecedented TON achieved by the Ni-BenzP*-phenoxide system, a


broadly applicable catalyst is expected to be less expensive and structurally less
exotic. In the same year as Limbach’s work (2014), Hendriksen et al.42 reported a
similar reaction, using stoichiometric LiI as a ‘‘hard’’ Lewis acid to form lithium
acrylate as the product. The inexpensive and readily available triethylamine and

Scheme 11. Limbach’s Catalytic System Consisting of Ni-dtbpe-tBuONa

218 Chem 3, 211–228, August 10, 2017


Scheme 12. Limbach’s Reaction with (BenzP*)nickelalactone as Intermediate

1,3-bis(dicyclohexylphosphino)propane (dcpp) were used as the optimal base and


ligand (Scheme 13). Based on the TON (21), it is less efficient than the Ni-BenzP*-
phenoxide system, but this could be partially due to the milder temperature
(50 C) than of Limbach’s reaction (100 C). Excessive zinc was still required as a facil-
itator, which precludes the wide utility of this method.

In 1991, Hoberg et al.43 found that the addition of a Lewis acid promoter such as
BeCl2 greatly facilitated ring contraction from 5-membered to 4-membered metal-
alactone, which readily underwent a second olefin insertion. Recently, Jin et al.44 re-
ported the distinct effect of sodium tetrakis[3,5-bis(trifluoromethyl)phenyl]borate
(NaBFAr4) with the diphosphine-ligated nickelalactone. It was found by DFT calcula-
tion (M06-L with LANL2DZ for Ni and 6-31G++(d,p) for other atoms) that ring
contraction of the five-membered nickelalactone (g-to-b isomerization) with sodium
coordination was much faster and almost 40 kJ/mol lower in activation energy than
the sodium-free reaction (Scheme 14). In a recent study, Plessow et al.45 also
concluded that the sodium cation might stabilize the five- and four-membered nick-
elalactone. However, it is believed that the use of Lewis acids will not necessarily
facilitate the formation of acrylic acid species unless a base is used.

For non-metal Lewis acids, Jin et al.46 reported the acceleration effect by tris(penta-
fluorophenyl)borane. Under ambient conditions, the borane was able to rapidly pro-
mote b-hydride elimination nickelalactone ligated with 1,10 -bis(diphenylphosphino)
ferrocene (dppf) to release the acrylate-borane salt (Scheme 15).

Molybdenum- and Tungsten-Mediated Reactions


For potential catalysts based on other transition metals, complexes based on Mo
and W have also garnered much interest. In 1985, Alvarez et al.47 first studied the
catalytic activities of Mo and W compounds with PMe2Ph or PMe3 as ligand. How-
ever, after the addition of CO2 and ethylene, the reaction halted upon the forma-
tion of the H-M-O(CO)R complex,48,49 with the acrylate as an anionic ligand on

Scheme 13. Vogt’s Reaction Facilitated by Lithium Salt

Chem 3, 211–228, August 10, 2017 219


Scheme 14. Stabilization Effect by Sodium Cation

Mo or W (Figure 2). Unfortunately, neither the monomeric (mostly seen in a W


complex) nor the dimeric species (Mo complex) led to product formation. In
2003, Schubert and Pápai50 performed DFT calculations on the PH3 and PMe3
ligated Mo complexes. This study indicated that the phosphine ligand plays a
crucial role in the reactivity of the complex. The overall transformation from the
Mo-ethylene-phosphine complex to the dimerized species was exothermic for
PMe3 but endothermic for PH3.

Interest in developing Mo and W catalysts has been regained recently: the Bern-
skoetter group at Brown University has reported a Mo complex based on the use
of the tridentate phosphine ligand.51 A zerovalent molybdenum pincer complex,
[(Ph2PCH2CH2)2PPh]Mo(C2H4) (N2)2, was found to facilitate the coupling of CO2
and ethylene to afford a molybdenum(II) acrylate hydride complex (Scheme 16).
Through kinetic isotope effect experiments, the substantial rate-determining step
was identified to be the oxidative addition of CO2 and ethylene to the Mo center,
whereas b-hydride elimination from the metallacylic complex and subsequent
dimerization were fast and not rate limiting. This observation was in strong contrast
with all Ni-, Ti-, Zr-, and Pd-catalyzed reactions, in which the rate-determining step
was the b-hydride elimination of metallalactone.

Zhang et al.52 later discovered that a Mo-tetrahydride complex was also capable of
inducing CO2-ethylene coupling. The structure was confirmed by NMR and X-ray
analysis. The formation of a new C–C bond of formate (formed with a mixed gas
rich in CO2), acrylate, and/or propionate (formed in the presence of H2) was
observed by in situ 1H and 31P NMR (Scheme 17).

In 2012, Wolfe and Bernskoetter53 discovered that the oxidative addition of CO2
and ethylene could be feasibly achieved at ambient temperature with the use
of trans-tetrakis(trimethylphosphite)tungsten bis(ethylene). This zero valent W
complex was prepared by sodium amalgam reduction of tungsten tetrachloride in
the presence of 2 bar of ethylene and 4 equiv of trimethylphosphite (Scheme 18).
The oxidative addition led to two isomeric intermediates k2-O,O and k3-C,C,O
with a rapid interconversion rate in solution. The structure and equilibration of the
two isomers were studied intensively by NMR analysis.

Eventually, the acrylate could be released from the W complex via treatment with
5 equiv of MeI to afford methyl acrylate in 31% yield (based per W) (Scheme 19).

F F
Ph Ph F B Ph Ph
P O
F F 3 P
Fe Ni Fe Ni
P O P OBFAr3
H
Ph Ph O Ph Ph

Scheme 15. Acceleration of b-hydride Elimination by an Electron-Deficient Borane

220 Chem 3, 211–228, August 10, 2017


O
PMe2Ph PMe2Ph
O
H P O
O
H Mo Mo
P W H

PhMe2P O PMe2Ph
O PMe2Ph

Figure 2. Resulting Complexes in the Early Study of Mo- and W-Assisted CO2-Ethylene Coupling
Reactions

Although this could prove to be a workable route to access acrylate, an efficient


release of the product is still underdeveloped.

In summary, reactions mediated by Mo and W complexes have less difficulty in b-hy-


dride elimination as a result of a different rate-determining step than in Ni catalysis.
However, only stoichiometric reactions were explored, and no catalytic example has
been realized.

Iron-Mediated Reactions
Iron is an economic and environmentally friendly catalyst source in comparison
with most other transition metals. However, potential catalytic ability of iron com-
plexes has not garnered wide exploration for CO2-ethylene coupling. In 1986,
Hoberg et al.54 reported the addition of butadiene and CO2 to the Fe complex,
in which a seven-membered ring was formed. Later, they found that the addi-
tion between CO2 and [(PEt3)2Fe(C2H4)2] could take place at a low temperature
of 78 C and under 1 bar of CO2.55 The resultant five-membered metallalactone
was able to undergo a variety of reactions depending on the ligand used. When
PMe3 was used, an Fe-acrylate complex was formed, which isomerized to four-
membered metallalactone and underwent a second CO2 insertion (Scheme 20).
Methylation of the dicarbonylated complex with the use of MeOH/HCl afforded
methyl methylmalonate. However, if dcpe (1,2-bis(dicyclohexylphosphino)ethane)
was used as the ligand, no ring contraction took place and dimethyl succinate
was produced.

In a recent study, Li et al.56 obtained the X-ray structure of (CH2 = CHCOOMe)


Fe(CO)4, and tested the reactivity of several Group VIII metal carbonyls toward
coupling with olefin and CO2. However, no desired reaction occurred.

Palladium-Mediated Reactions
Palladium complexes have also been explored as a potential class of catalyst for
CO2-ethylene coupling. Reactions of CO2 with other olefinic compounds such as

Scheme 16. The Expected Catalytic Cycle of Bernskoetter’s Mo catalysis

Chem 3, 211–228, August 10, 2017 221


O
C
Ph H H Ph O
P PPh2 P PPh2 Ph
C2H4 (4 bar) CO2 P PPh2
Mo Mo
P PPh3 Mo
P
Ph2 H Ph2 P
H Ph2

Me

Ph O O O O
Ph
PPh3 P PPh2 H2 P PPh2
Mo Mo
P PPh3 P PPh3
Ph2 Ph2
H H
2-acrylate 2-propionate

Scheme 17. Formation of Carbonylation Products by Bernskoetter’s Mo-tetrahydride Complex

dienes were investigated by several groups over the years. The earliest examples
were presented in a series of studies by several groups, focusing on the reaction
of butadiene. In 1976, Sasaki et al.57 reported a CO2-butadiene addition catalyzed
by 0.015 mol % of Pd(dppe) complex. The major product was (E)-octa-1,3,7-triene,
however, up to 12% yield of a g-lactone (2-ethylidenehept-5-en-4-olide) was ob-
tained, indicating that CO2 insertion had taken place. Musco et al.58 also studied
the Pd-mediated CO2-butadiene addition, which formed a complex set of products
with multiple butadiene fragments. Later in 1986, Behr et al.59 reported that two
molecules of butadiene reacted with CO2 to form a d-lactone (2-ethylidene-6-
hepten-5-olide), in the presence of a PiPr3-ligated Pd catalyst. Behr60 also found
that the presence of nitrile co-ligands was crucial to create vacant coordination sites,
thus forming the catalytically active species. In recent years, Behr’s group has been
transferring the CO2-butadiene addition reaction to a miniplant with a continuous
setup.61 As for substituted dienes, Hoberg and Minato62 reported the dimerization
of isoprene with one molecule of CO2 to form unsaturated esters with the assistance
of a stoichiometric amount of tributyltin ethoxide.

Scheme 18. Preparation of a W complex and the Formation-Isomerization of the Oxidative


Addition Product

222 Chem 3, 211–228, August 10, 2017


Scheme 19. Release of Methyl Acrylate from the W complex

The activity of Pd complexes in catalyzing monoolefin carboxylation was investi-


gated by Osakada et al.63 and Yamamoto et al.64 With a zero valent Pd coordi-
nated with tertiary phosphine ligands, the coupling of CO and 3-butenoic
acid went smoothly to give methylsuccinic anhydride and glutaric anhydride.
Although this study was not directly related to acrylate synthesis, it aided in the
understanding of the properties and reactivities of five- and six-membered pallada-
lactone (Scheme 21).

In 2007, Aresta et al.65 reported a reaction of pre-formed Pd–COOMe complex


with ethylene and CO2, for synthesizing ethyl acrylate. In the proposed catalytic cy-
cle (although being stoichiometric in the actual reaction), the catalytically active
[(dppe)PdH]+ readily reacted with ethylene and formed the [(dppe)PdEt]+ complex,
which underwent CO2 and ethylene insertions and eventually gave ethyl acrylate
in free form (Scheme 22). In contrast to other catalytic systems, this creative
method involved no metallalactone, thus the b-hydride elimination in the acyclic
structure was quite feasible. However, the proposed mechanism was oversimpli-
fied, because it involved a less likely non-traditional ‘‘inverse-CO2 insertion,’’ in
which the energy barrier was as high as 58 kcal/mol (243 kJ/mol). In summary,
there are two major limitations of this reaction: (1) the requirement for a nearly stoi-
chiometric amount of Pd complex and (2) no esters other than ethyl acrylate could
be synthesized.

Recently, Stieber et al.66 developed a Pd catalyst similar to their Ni-BenzP*-phen-


oxide catalyst. Like the Ni system, sodium 2-F-phenoxide (2-F–PhONa) was used
as the optimal base (Scheme 23). Again, the addition of zinc powder greatly facili-
tated the reaction. Zinc was believed to initiate the reaction by reducing Pd(II) to
Pd(0). Bidentate ligand 1,2-bis(dicyclohexylphosphino)-ethane (dcpe) was identified
as the optimal ligand. At 145 C, with 10 bar of ethylene and 20 bar of CO2, a TON of

O
L = DCPE MeO
L O MeOH
Fe O O
O L Fe
L CO2 HCl O
L O
OMe
O
L = PMe3

O O
Me O MeOH MeO
L H L CO2 L
Fe Me Me
L
Fe
O O L Fe L
O HCl MeO
O O
O O

Scheme 20. Fe-Mediated Transformations Leading to Multiple Products

Chem 3, 211–228, August 10, 2017 223


Scheme 21. Pd-Catalyzed Methylsuccinic Anhydride and Glutaric Anhydride Synthesis by
Osakada and Yamamoto

29 was realized. Note that the pressure of each gaseous component was doubled
from that in the Ni(BenzP*) system. The previously used ligand dtbpe afforded a
much lower TON (1.5).

With the same catalytic system, Stieber et al.66 were able to couple CO2 with dienes
to afford the corresponding conjugated acrylate analogs with TONs of 24–50
(Scheme 24).

Platinum- and Rhodium-Mediated Reactions


There have been only a few reports on the chemical properties of platinumalactone
complexes. However, the complexes were either not formed by CO2-ethylene addi-
tion67 or other olefins were required instead of ethylene.68,69 No subsequent release
of acrylate was reported in these studies.

For rhodium-mediated reactions, there was a sole example reported by Aresta


and Quaranta70 whereby [Rh(bpy) (C2H4)Cl] was exposed to an atmosphere of
CO2 (1 bar) for 9 days. The corresponding rhodialactone was isolated and confirmed
by NMR and IR spectroscopy (Scheme 25). One might be skeptical about the reac-
tivity of the complex, but it is worth noting that the pressure of CO2 is much lower
than in other metal-mediated reactions, and the formation process was under

Scheme 22. Proposed Mechanism of Dibenedetto and Pápai’s Ethyl Acrylate Synthesis Catalyzed
by the (dppe)Pd–H Complex

224 Chem 3, 211–228, August 10, 2017


Scheme 23. Limbach’s Pd(dcpe)-Phenoxide Catalytic System

room temperature. There were no reported subsequent transformations of this rho-


dialactone, and studies on this complex have been largely unexplored.

CONCLUSIONS AND OUTLOOK


Metal-catalyzed CO2-ethylene coupling to synthesize acrylic acid derivatives is
an ideal reaction in the financially and environmentally conscious future. Over the
past four decades, tremendous efforts have been made to develop an effective cat-
alytic system, and this continues to garner much attention.

Among the transition-metal complexes with catalytic activities, nickel complexes are
still the most deeply investigated potential catalysts. Mo- and W-mediated reactions
have a unique rate-determining step, so that the difficulty of b-hydride elimination
can be avoided. However, current examples are not catalytic and are at the stage
of mechanistic study. Pd catalysis, exemplified by the ethyl acrylate synthesis
catalyzed by the Pd–H complex, enables an alternative approach and warrants
further investigation. Reactions induced by Fe, Pt, and Rh complexes are currently
underdeveloped.

In conclusion, the major limitations of existing methods are as follows:

1. To date, no catalytic system has been efficient enough for industrial practice as
a result of the extremely low TON.
2. Although the efficacy of several catalytic systems has been demonstrated,
the requirement for a superstoichiometric amount of a reductant (e.g., zinc)
and/or an alkylation regent (e.g., MeI) highlights the challenges of this field.

Scheme 24. Coupling of Dienes with CO2 Catalyzed by the Pd(dcpe) Complex

Chem 3, 211–228, August 10, 2017 225


N 9 days N
Rh + CO2 Rh
Cl THF, rt O O
N N
(1 bar)

Scheme 25. Aresta’s Rhodialactone Formation Reaction

Even in the theoretical study, the unclear function of zinc and the variable
mechanism of methylation by MeI are still puzzles to solve.
3. Current methods produce esters and sodium salts of acrylic acid, which
are extremely valuable fine chemicals among their class. Unfortunately, no
example to date has been able to produce free acrylic acid. Therefore, it is a major
disadvantage in comparison with the traditional propylene oxidation process.
4. Strong or structurally exotic bases are often necessary, which is neither consid-
ered safe nor economically attractive.
5. Most examples require harsh reaction conditions such as high temperature
and pressure.

For the establishment of a more efficient catalytic process, ligand design is believed
by previous researchers and by the authors of this review to be the key to success
because it is the most straightforward way to modify the catalytic property of the
metal complex. We also envision the following potential solutions:

1. An amphoteric compound with both a Lewis acid part and Brønsted base part
could have the ability to stabilize carbonyl and induce ring contraction, thus
liberating free acrylic acid (instead of sodium acrylate). The acrylic acid can
either be removed from the system by separation techniques or be neutralized
with a mild and cheap base.
2. For feasible b-hydride elimination of the metalalactone, adding a methyl-
ation reagent is a popular strategy, which however often introduces unde-
gradable contaminanta such as iodide. An alternative solution would be to
use a cleaner and traceless reagent such as dimethyl carbonate (DMC).
Ideally, DMC can be generated from CO2 and methanol in situ by an orthog-
onal catalytic system, which also catalyzes the formation of acrylate in the
same reactor.
3. The inherent geometric limitation for b-hydride elimination of a 5-membered
metalalactone is one of the biggest obstacles in the existing examples. An
intriguing strategy is to avoid the formation of metalalactone. To that end,
we expect novel C–H activation methods to be developed that directly acti-
vate the C–H bond of ethylene and insert the vinyl group to CO2, thus the
problem with b-hydride elimination can be skipped.

Despite the formidable challenges, recent progress in catalytic CO2-ethylene addi-


tion is encouraging and inspirational. In addition to realizing the dream reaction for
producing acrylate in a low-carbon and sustainable manner, these recent advances
are important for developing an understanding and improving other types of CCU
transformations that involve the reduction of CO2.

AUTHOR CONTRIBUTIONS
X.W. proposed the topic of the review and conceptualized the article content. X.W.,
H.W., and Y.S. discussed and finalized the manuscript.

226 Chem 3, 211–228, August 10, 2017


ACKNOWLEDGMENTS
This work is supported by China’s Ministry of Science and Technology under contract
number 2017YFB0602200. We thank Dr. Andrew T. Parsons (Amgen) for his help in
preparing this manuscript.

REFERENCES AND NOTES


1. Ohara, T., Sato, T., Shimizu, N., Prescher, G., molecular solution to a global challenge? 25. Graham, D.C., Mitchell, C., Bruce, M.I., Metha,
Schwind, H., Weiberg, O., Marten, K., and Angew. Chem. Int. Ed. 50, 8510–8537. G.F., Bowie, J.H., and Buntine, M.A. (2007).
Greim, H. (2003). Acrylic acid and derivatives. In Production of acrylic acid through nickel-
Ullmann’s Encyclopedia of Industrial 14. Gomes, C.D.N., Jacquet, O., Villiers, C., Thury, mediated coupling of ethylene and carbon
Chemistry, B. Elvers, ed. (Wiley-VCH Verlag). P., Ephritikhine, M., and Cantat, T. (2012). A dioxide – a DFT study. Organometallics 26,
http://dx.doi.org/10.1002/14356007.a01_161. diagonal approach to chemical recycling of 6784–6792.
pub3. carbon dioxide: organocatalytic transformation
for the reductive functionalization of CO2. 26. Pápai, I., Schubert, G., Mayer, I., Besenyei, G.,
2. Mori, H., and Müller, A.H.E. (2003). New Angew. Chem. Int. Ed. 51, 187–190. and Aresta, M. (2004). Mechanistic details of
polymeric architectures with (meth)acrylic acid nickel(0)-assisted oxidative coupling of CO2
segments. Prog. Polym. Sci. 28, 1403–1439. 15. Liu, Q., Wu, L., Jackstell, R., and Beller, M. with C2H4. Organometallics 23, 5252–5259.
(2014). Using carbon dioxide as a building
3. Teles, J.H., Hermans, I., Franz, G., and Sheldon, block in organic synthesis. Nat. Commun. 6, 27. Hoberg, H., Peres, Y., and Milchereit, A. (1986).
R.A. (2015). Oxidation. In Ullmann’s 5933. C–C-Verknüpfung von Alkenen mit CO2 an
Encyclopedia of Industrial ChemistryB. Elvers, Nickel(0); Herstellung von Zimtsäure aus Styrol.
ed. (Wiley-VCH Verlag). http://dx.doi.org/10. 16. Yu, B., Diao, Z.-F., Guo, C.-X., and He, L.-N. J. Organomet. Chem. 307, C38–C40.
1002/14356007.a18_261. (2013). Carboxylation of olefins/alkynes with 28. Fischer, R., Langer, J., Görls, H., Malassa, A.,
CO2 to industrially relevant acrylic acid Walther, D., and Vaughan, G. (2006). A key step
4. Campbell, W.E., McDaniel, E.L., Reece, W.H.,
derivatives. J. CO2 Util 1, 60–68. in the formation of acrylic acid from CO2 and
Williams, J.E., and Young, H.S. (1970).
Oxidation of propylene to acrylic acid over a ethylene: the transformation of a
17. Juliá-Hernández, F., Gaydou, M., Serrano, E., nickelalactone into a nickel-acrylate complex.
catalyst containing oxides of arsenic, niobium,
Van Gemmeren, M., and Martin, R. (2016). Chem. Commun. 23, 2510–2512.
and molybdenum. Ind. Eng. Chem. Prod. Res.
Ni- and Fe-catalyzed carboxylation of
Dev. 9, 325–334.
unsaturated hydrocarbons with CO2. Top. Curr. 29. Walther, D., Dinjus, E., Sieler, J., Andersen, L.,
5. Landi, G., Lisi, L., and Russo, G. (2005). Chem. (J) 374, 45, 1–38. and Lindqvist, O. (1984). Aktivierung von
Oxidation of propane and propylene to acrylic Kohlendioxid an Übergangsmetallzentren:
acid over vanadyl pyrophosphate. J. Mol. 18. Hoberg, H., and Schaefer, D. (1982). Nickel(0)- Metallaringschluss mit Dicyclopentadien am
Catal. A Chem. 239, 172–179. induzierte CC-verknüpfung zwischen alkenen elektronenreichen Nickel(0)-komplexrumpf als
und kohlendioxid. J. Organomet. Chem. 236, topo- und stereoselektive Reaktion.
6. Samel, U.-R., Kohler, W., Gamer, A.O., Keuser, C28–C30. J. Organomet. Chem. 276, 99–107.
U., Yang, S.-T., Jin, Y., Lin, M., and Wang, Z.-Q.
(2014). Propionic acid and derivatives. In 19. Hoberg, H., Schaefer, D., Burkhart, G., Kruger, 30. Walther, D., Bräunlich, G., Kempe, R., and
Ullmann’s Encyclopedia of Industrial C., and Romao, M. (1984). Nickel(0)-induzierte Sieler, J. (1992). Aktivierung von CO2 an
Chemistry, B. Elvers, ed. (Wiley-VCH Verlag). CC-verknüpfung zwischen kohlendioxid und Übergangsmetallzentren: Zum Ablauf der
http://dx.doi.org/10.1002/14356007.a22_223. alkinen sowie alkenen. J. Organomet. Chem. homogen-katalytischen Bildung von 2-Pyron
pub2. 266, 203–224. aus Kohlendioxid und hex-3-in an Nickel(0)-
fragmenten. J. Organomet. Chem. 436,
7. Brooks, R.E., Gresham, W.F., Hardy, J.V.E., and 20. Hoberg, H., Ballesteros, A., Sigan, A., Jegat, C., 109–119.
Lupton, J.M. (1957). Synthesis of propionic and Milchereit, A. (1991). Durch (lig)Ni(0)
anhydride and propionic acid. Ind. Eng. Chem. induzierte Herstellung von mono- und 31. Langer, J., Fischer, R., Görls, H., and Walther,
49, 2004–2007. di-Carbonsäuren aus Cyclopenten und D. (2004). A new set of nickelacyclic
Kohlendioxid. Synthesis 1991, 395–398. carboxylates containing pyridine as supporting
8. Alper, H., Woell, J.B., Despeyroux, B., and ligand: synthesis, structures and application
Smith, D.J.H. (1983). The regiospecific 21. Walther, D., and Dinjus, E. (1982). Aktivierung in CC- and CS- linkage reactions.
palladium catalysed hydrocarboxylation of von Kohlendioxid an Übergangsmetallzentren; J. Organomet. Chem. 689, 2952–2962.
alkenes under mild conditions. J. Chem. Soc. Die metallaringschlußreaktion zwischen
Chem. Commun. 1270–1271. Kohlendioxid und 1,3-Dienen am 32. Bruckmeier, C., Lehenmeier, M.W., Reichardt,
elektronenreichen Nickel(0)-Komplexrumpf. R., Vagin, S., and Rieger, B. (2010). Formation of
9. Goedheijt, M.S., Reek, J.N.H., Kamer, P.C.J., Z. Chem. 22, 228–229. methyl acrylate from CO2 and ethylene via
and van Leeuwen, P.W.N.M. (1998). A highly methylation of nickelalactones.
selective water-soluble dicationic palladium 22. Walther, D., Dinjus, E., Sieler, J., Thanh, N.N., Organometallics 29, 2199–2202.
catalyst for the biphasic hydroxycarbonylation and Schade, W. (1983). Aktivierung von CO2 an
of alkenes. Chem. Commun. 2431–2432. 33. Bräunlich, G., Fischer, R., Nestler, B., and
Übergangsmetallzentren: Struktur und Walther, D. (1989). Metallacyclische
Reaktivität eines C–C-Kopplungsproduktes Carboxylate des Nickels: Synthone zur C–C-
10. Reppe, W., and Kröper, H. 1952. German
von CO2 und 2.3-Dimethylbutadien am Verknüpfung via Kreuzkopplung. Z. Chem. 29,
patent 848,355.
elektronenreichen Nickel(0). Z. Naturforsch. 417.
11. Reppe, W., and Stadler, R. 1962. Process for 38B, 835–840.
producing acrylic acid. US patent 3023237, filed 34. Schönecker, B., Walther, D., Fischer, R.,
May 1, 1959, and published February 27, 1962. 23. Hoberg, H., and Schaefer, D. (1983). Nickel(0)- Nestler, B., Bräunlich, G., Eibisch, H., and
https://www.google.com/patents/US3023237. induzierte CC-Verknüpfung zwischen Droescher, P. (1990). A nickelacycle as
Kohlendioxid und Ethylen sowie mono- oder propionic acid equivalent for carbon-carbon
12. Chafetz, H., and Patterson, J.A. 1966. Carbonyl di-substituierten Alkenen. J. Organomet. coupling reactions: application to the synthesis
process. US patent 3282993, filed February 19, Chem. 251, C51–C53. of C25 steroid carboxylic acids. Tetrahedron
1962, and published November 1, 1966. Lett. 31, 1257–1260.
24. Hoberg, H., Peres, Y., Krüger, C., and Tsay,
13. Cokoja, M., Bruckmeier, C., Rieger, B., Y.-H. (1987). A 1-oxa-2-nickela-5- 35. Fischer, R., Walther, D., Bräunlich, G.,
Herrmann, W.A., and Kühn, F.E. (2011). cyclopentanone from ethene and carbon Undeutsch, B., and Ludwig, W. (1992).
Transformation of carbon dioxide with dioxide: preparation, structure, and reactivity. Nickelalactone als Synthesebausteine:
homogeneous transition-metal catalysts: a Angew. Chem. Int. Ed. 26, 771–773. Sonochemische und Bimetallaktivierung der

Chem 3, 211–228, August 10, 2017 227


Kreuzkopplungsreaktion mit Alkyl- carbon dioxide with ethylene complexes of von 1,3-Dienen mit Kohlendioxid. Chem. Ber
halogeniden. J. Organomet. Chem. 427, molybdenum and tungsten. J. Am. Chem. Soc. 119, 991–1015.
395–407. 107, 5529–5531.
60. Behr, A. (1988). Carbon Dioxide Activation by
36. Castaño, A.M., and Echavarren, A.M. (1994). 48. Alvarez, R., Carmona, E., Galindo, A., Gutiérrez, Metal Complexes (VCH).
Reactivity of a nickelacycle derived from E., Marı́n, J.M., Monge, A., Poveda, M.L., Ruiz,
aspartic acid: alkylations, insertions, and C., and Savariault, J.M. (1989). Formation of 61. Behr, A., and Becker, M. (2006). The
oxidations. Organometallics 13, 2262–2268. carboxylate complexes from the reactions of telomerisation of 1,3-butadiene and carbon
COP with ethylene complexes of molybdenum dioxide: process development and
37. Guo, W., Michel, C., Schwiedernoch, R., and tungsten. X-ray and neutron diffraction optimisation in a continuous miniplant. Dalton
Wischert, R., Xu, X., and Sautet, P. (2014). studies. Organometallics 8, 2430–2439. Trans. 4607–4613.
Formation of acrylates from ethylene and CO2
on Ni complexes: a mechanistic viewpoint from 49. Galindo, A., Pastor, A., Pérez, P.J., and
62. Hoberg, H., and Minato, M. (1991). Palladium-
a hybrid DFT approach. Organometallics 33, Carmona, E. (1993). Bis(ethy1ene) complexes
catalysed dimerization of isoprene with carbon
6369–6380. of molybdenum and tungsten and their
dioxide in the presence of organotin ethoxide
reactivity toward CO2. New examples of
38. Lee, S.Y.T., Cokoja, M., Drees, M., Li, Y., Mink, and 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU).
acrylate formation by coupling of ethylene and
J., Herrmann, W.A., and Kühn, F.E. (2011). J. Organomet. Chem. 406, C25–C28.
carbon dioxide. Organometallics 12, 4443–
Transformation of nickelalactones to methyl 4451.
acrylate: on the way to a catalytic conversion of 63. Osakada, K., Doh, M.-K., Ozawa, F., and
carbon dioxide. ChemSusChem 4, 1275–1279. 50. Schubert, G., and Pápai, I. (2003). Acrylate Yamamoto, A. (1990). Catalytic and
formation via metal-assisted CC coupling stoichiometric carbonylation of b,g-
39. Plessow, P.N., Weigel, L., Lindner, R., Schäfer, between CO2 and C2H4: reaction mechanism unsaturated carboxylic acids to give cyclic
A., Rominger, F., Limbach, M., and Hofmann, P. as revealed from density functional anhydrides through intermediate palladium-
(2013). Mechanistic details of the nickel- calculations. J. Am. Chem. Soc. 125, 14847– containing cyclic esters. Organometallics 9,
mediated formation of acrylates from CO2, 14858. 2197–2198.
ethylene and methyl iodide. Organometallics
32, 3327–3338. 51. Bernskoetter, W.H., and Tyler, B.T. (2011). 64. Yamamoto, T., Sano, K., Osakada, K., Komiya,
Kinetics and mechanism of molybdenum- S., Yamamoto, A., Kushi, Y., and Tada, T. (1990).
40. Lejkowski, M.L., Lindner, R., Kageyama, T., mediated acrylate formation from carbon Comparative studies on reactions of a,b- and
Bódizs, G.É., Plessow, P.N., Müller, I.B., dioxide and ethylene. Organometallics 30, b,g-unsaturated amides and acids with
Schäfer, A., Rominger, F., Hofmann, P., Futter, 520–527. nickel(0), palladium(0), and platinum(0)
C., et al. (2012). The first catalytic synthesis of an complexes. Preparation of new five- and six-
acrylate from CO2 and an alkene  a rational 52. Zhang, Y., Hanna, B.S., Dineen, A., Williard, membered nickel- and palladium-containing
approach. Chem. Eur. J. 18, 14017–14025. P.G., and Bernskoetter, W.H. (2013). cyclic amide and ester complexes.
Functionalization of carbon dioxide with Organometallics 9, 2396–2403.
41. Huguet, N., Jevtovikj, I., Gordillo, A., Lejkowski, ethylene at molybdenum hydride complexes.
M.L., Lindner, R., Bru, M., Khalimon, A.Y., Organometallics 32, 3969–3979. 65. Aresta, M., Pastore, C., Giannoccaro, P.,
Rominger, F., Schunk, S.A., Hofmann, P., and
Kovács, G., Dibenedetto, A., and Pápai, I.
Limbach, M. (2014). Nickel-catalyzed direct 53. Wolfe, J.M., and Bernskoetter, W.H. (2012).
(2007). Evidence for spontaneous release of
carboxylation of olefins with CO2: one-pot Reductive functionalization of carbon dioxide
acrylates from a transition-metal complex upon
synthesis of a,b-unsaturated carboxylic acid to methyl acrylate at zerovalent tungsten.
coupling ethene or propene with a carboxylic
salts. Chem. Eur. J. 20, 16858–16862. Dalton Trans. 41, 10763–10768.
moiety or CO2. Chem. Eur. J. 13, 9028–9034.
42. Hendriksen, C., Pidko, E.A., Yang, G., 54. Hoberg, H., Jenni, K., Krüger, C., and Raabe, E.
Schäffner, B., and Vogt, D. (2014). Catalytic (1986). CC coupling of CO2 and butadiene on 66. Stieber, S.C.E., Huguet, N., Kageyama, T.,
formation of acrylate from carbon dioxide and iron(0) complexes  a novel route to a,u- Jevtovikj, I., Ariyananda, P., Gordillo, A.,
ethene. Chem. Eur. J. 20, 12037–12040. dicarboxylic acids. Angew. Chem. Int. Ed. 25, Schunk, S.A., Rominger, F., Hofmann, P., and
810–811. Limbach, M. (2015). Acrylate formation from
43. Hoberg, H., Ballesteros, A., Sigan, A., Jegat, C., CO2 and ethylene: catalysis with palladium and
Bärhausen, D., and Milchereit, A. (1991). 55. Hoberg, H., Jenni, K., Angermund, K., and mechanistic insight. Chem. Commun. 51,
Ligandgesteuerte Ringkontraktion von Krüger, C. (1987). CC-linkages of ethene with 10907–10909.
Nickela-fünf-in Vierringkomplexe-neuartige CO2 on an iron(0) complex  synthesis and
Startsysteme für die präparative Chemie. crystal structure analysis of [(PEt3)2Fe(C2H4)2]. 67. Aye, K.-T., Colpitts, D., Ferguson, G., and
J. Organomet. Chem. 407, C23–C29. Angew. Chem. Int. Ed. 26, 153–155. Puddephatt, R.J. (1988). Activation of a
b-lactone by oxidative addition and the
44. Jin, D., Williard, P.G., Hazari, N., and 56. Li, B., Kyran, S.J., Yeung, A.D., Bengali, A.A., structure of a platina(IV)lactone.
Bernskoetter, W.H. (2014). Effect of sodium and Darensbourg, D.J. (2013). Acrylic acid Organometallics 7, 1454–1456.
cation on metallacycle b-hydride elimination in derivatives of group 8 metal carbonyls: a
CO2ethylene coupling to acrylates. Chem. structural and kinetic study. Inorg. Chem. 52, 68. Sano, K., Yamamoto, T., and Yamamoto, A.
Eur. J. 20, 3205–3211. 5438–5447. (1984). Preparation of Ni- or Pt-containing
cyclic esters by oxidative addition of cyclic
45. Plessow, P.N., Schäfer, A., Limbach, M., and 57. Sasaki, Y., Inoue, Y., and Hashimoto, H. (1976).
carboxylic anhydride and their properties. Bull.
Hofmann, P. (2014). Acrylate formation Reaction of carbon dioxide with butadiene
Chem. Soc. Jpn. 57, 2741–2747.
from CO2 and ethylene mediated by catalysed by palladium complexes. Synthesis
nickel complexes: a theoretical study. of 2-ethylidenehept-5-en-4-olide. J. Chem.
69. Schulz, P.S., Walter, O., and Dinjus, E. (2005).
Organometallics 33, 3657–3668. Soc. Chem. Commun. 605–606.
Facile synthesis of a tricyclohexylphosphine-
46. Jin, D., Schmeier, T.J., Williard, P.G., Hazari, N., 58. Musco, A., Perego, C., and Tartiari, V. (1978). stabilized h3-allyl-carboxylato Ni(II) complex
and Bernskoetter, W.H. (2013). Lewis acid Telomerization reactions of butadiene and and its relevance in electrochemical butadiene
induced b-elimination from a nickelalactone: CO2 catalyzed by phosphine Pd(0) complexes: carbon dioxide coupling. Appl. Organomet.
efforts toward acrylate production from CO2 (E)-2-ethylidenehept-6-en-5-olide and Chem. 19, 1176–1179.
and ethylene. Organometallics 32, 2152–2159. octadienyl esters of 2-ethylidenehepta-4,6-
dienoic acid. Inorg. Chim. Acta 28, L147–L148. 70. Aresta, M., and Quaranta, E. (1993).
47. Alvarez, R., Carmona, E., Cole-Hamilton, D.J., Synthesis, characterization and reactivity of
Galindo, A., Gutierrez-Puebla, E., Monge, A., 59. Behr, A., He, R., Juszak, K.D., Krüger, C., and [Rh(bpy)(C2H4)Cl]. A study on the reaction with
Poveda, M.L., and Ruiz, C. (1985). Formation Tsay, Y. (1986). Steuerungsmöglichkeiten bei C1 molecules (CH2O, CO2) and NaBPh4.
of acrylic acid derivatives from the reaction of der Übergangsmetall-katalysierten umsetzung J. Organomet. Chem. 463, 215–221.

228 Chem 3, 211–228, August 10, 2017

You might also like