Anda di halaman 1dari 10

G Model

JCOU-6; No. of Pages 10


Journal of CO2 Utilization xxx (2013) xxxxxx

Contents lists available at SciVerse ScienceDirect

Journal of CO2 Utilization


journal homepage: www.elsevier.com/locate/jcou

Review

Thermal, electrochemical, and photochemical conversion of CO2 to fuels and


value-added products
Boxun Hu a, Curtis Guild a, Steven L. Suib a,b,*
a
b

Department of Chemistry, University of Connecticut, Unit 3060, 55 North Eagleville Road, Storrs, CT 06269-3060, USA
Institute of Materials Science, University of Connecticut, USA

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 27 December 2012
Received in revised form 14 March 2013
Accepted 15 March 2013

This review compares various alternate fuels and value-added products from conversion of carbon
dioxide such as simple molecules to higher hydrocarbon fuels and polymers. Different methods of
activation are summarized that lead to different products. We summarize the advantages and
disadvantages of different methods of conversion of carbon dioxide. An overall summary is given at the
end of the review that discusses future approaches and promising approaches.
2013 Elsevier Ltd. All rights reserved.

Keywords:
Review
Carbon dioxide
Activation

Contents
1.
2.
3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Thermodynamics and kinetics of CO2 conversion . . . . . . . . . . . . . . . . . . .
CO2 conversion to fuel and value-added products . . . . . . . . . . . . . . . . . . .
CO2 conversion to CO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
CO production by reduction of CO2 . . . . . . . . . . . . . . . . .
3.1.1.
CO production by electrocatalysis. . . . . . . . . . . . . . . . . . .
3.1.2.
CO production by plasma . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.3.
Synthesis gas production by reforming . . . . . . . . . . . . . .
3.1.4.
CO2 conversion to HCOOH and HCHO . . . . . . . . . . . . . . . . . . . . . . .
3.2.
CO2 conversion to CH3OH. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.
CO2 conversion to long chain hydrocarbons and oxygenates . . . . .
3.4.
CO2 as building blocks for oxygen-rich compounds and polymers
3.5.
Prospective in CO2 conversion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
Carbon dioxide (CO2) utilization technologies have emerged to
reduce CO2 emissions by developing benecial uses of CO2 [14]. In
todays world, two major environmental concerns attributed in

* Corresponding author at: Department of Chemistry, University of Connecticut,


Unit 3060, 55 North Eagleville Road, Storrs, CT 06269-3060, USA.
Tel.: +1 860 486 2797; fax: +1 860 486 2981.
E-mail address: steven.suib@uconn.edu (S.L. Suib).

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

000
000
000
000
000
000
000
000
000
000
000
000
000
000
000

part to buildup of carbon dioxide (CO2), the acidication of the


oceans and global warming [5,6]. These issues are consequences in
part to atmospheric CO2 concentrations (ACC) rising around the
globe. Fossil fuel consumption has caused ACC increases from 280
parts per million (ppm) in pre-industrial times to 382 ppm in 2006
according to the National Oceanic and Atmospheric Administration (NOAA). Currently, the atmospheric CO2 concentration is still
steadily increasing at a rate of about 1.9 ppm/year. In tandem with
this, the international energy outlook (2011) has projected that
world energy consumption will increase 53 percent from 2008 to
2035 [7]. Scientists from the Intergovernmental Panel on Climate

2212-9820/$ see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jcou.2013.03.004

Please cite this article in press as: B. Hu, et al., J. CO2 Util. (2013), http://dx.doi.org/10.1016/j.jcou.2013.03.004

G Model

JCOU-6; No. of Pages 10


B. Hu et al. / Journal of CO2 Utilization xxx (2013) xxxxxx

Change (IPCC) have suggested that 350 ppm CO2 is likely the safe
upper limit for atmospheric greenhouse gases [8]. To realize this
goal, CO2 must be captured and either stored or converted into
global warming neutral impact compounds. Utilization of CO2
provides an attractive avenue for this objective.
CO2 is extensively used for enhanced oil recovery, urea and
polymer synthesis as a monomer feedstock, the food and beverage
industry as a propellant, and chemical production [9]. However,
only less than 1% of global anthropogenic CO2 generated are
utilized to these ends. The rest is released to the atmosphere due to
lack of economical technologies to convert these C1 sources to
commodity products, and therein lack of demand. CO2 conversion
to fuel and value-added products is an ideal route for CO2
utilization due to the simultaneous disposal of CO2 and the benet
that many products can be used as alternate transportation fuels.
CO2 is a kinetically and thermodynamically stable molecule,
thus CO2 conversion reactions are endothermic and need efcient
catalysts to obtain high yield. Various renewable energies such as
solar, wind, and hydroelectric are proposed as energy sources for
CO2 conversion. These largely intermittent kinetic energies are
stored in the alternate fuels in the stable form of chemical energy
that can be transported and used on demand. These synthetic fuels
will sustain new and expanding markets in transportation.
Furthermore, the synthetic fuels produced by CO2 utilization are
compatible with current hydrocarbon-based automobiles and
transportation systems. The products of carbon dioxide conversion
can supplement or replace chemical feedstocks in the chemical,
pharmaceutical, and polymer industries.
This review summarizes the various alternate fuels and valueadded products from CO2 conversion in order of products from
simple molecules to higher hydrocarbon fuels and polymers. Each
product may be produced by several methods. The pros and cons
of different conversion approaches are compared, with the intent
of giving direction to the reader in selecting a proper approach to
meet specic needs. Many good reviews and books [3,10] have
described CO2 utilization by photocatalytic synthesis [4,1113],
electrochemical reduction [14,15], plasma [16], and other
methods [1,17,18]. We do not attempt to duplicate these
references but summarize possible reaction mechanisms, unique
catalyst and experimental design, structure-activity relationships, and energy efciency. These issues provide help in
understanding the product selectivity and catalytic activity of
each system, and evaluate the potential practical applications of
the process.
2. Thermodynamics and kinetics of CO2 conversion
Fig. 1 gives the Gibbs free energy of CO2 and the products
converted from CO2 (CRC Handbook of Chemistry and Physics)
[19]. CO2 molecules have a highly stable linear and centrosymmetric (O5
5C5
5O) structure. The difference in Gibbs free energy
(DG) between the product and reactants at specic reaction is the
driving force shown in Eq. (1):

DG DH T DS

(1)

where DH is the enthalpy change, DS is the entropy change, and T is


temperature.
Most conversions of CO2 to these listed products are
endothermic reactions like natural photosynthesis as shown in
Eq. (2):
6CO2 g 6H2 Ol hn ! C6 H12 O6 s 6O2 g

(2)

There are several kinds of driving forces for the CO2 conversion.
First, highly efcient articial enzymes lower the activation energy
required for photoreduction of CO2. These exhibit rates faster than

Fig. 1. Gibbs free energy of CO2 and its related products.

in nature, though the cost of the catalysts is still a concern.


Secondly, inorganic nanocatalysts promote conversion to valueadded olens in the FischerTropsch synthesis. Third, while not
catalytic, applied potentials serve to facilitate the electrochemical
reduction of CO2. Fourth, CO2 will react with other reactants with
higher Gibbs free energy, such as methane and hydrogen. Last but
not least, CO2 reacts with other monomeric species to form
polymers. These polymerization reactions run at mild conditions
with high turnover numbers. In a practical operation, two or more
types of principles may function simultaneously and make the CO2
conversion more efcient. The net result of these processes is the
conversion of different types of energies into chemical energy.
3. CO2 conversion to fuel and value-added products
3.1. CO2 conversion to CO
CO2 conversion to CO (carbon monoxide) looks like the simplest
route for CO2 reduction. CO is a feedstock or intermediate for the
production of methanol and hydrocarbon fuels via the Fischer
Tropsch synthesis [20]. Several technical routes lead to the CO2
conversion to CO, such as photoreduction, electrolysis, plasma,
electrocatalysis, dry- and bi-reforming, and (tri-) reforming.
Different types of catalysts are involved in these processes.
3.1.1. CO production by reduction of CO2
Mimicking natures photosynthesis process, photoreduction of
CO2 is one of the most alluring methods for CO2 conversion due to
the abundance and free access of sunlight. To meet with global
energy demands, Lewis and Nocera have proposed to convert and
store solar energy in chemicals (H2, methanol, and hydrocarbons)
via the photosynthetic process [21]. Solar radiation varies across
the globe from altitude, height, atmospheric conditions, and
season. For example, solar radiation for at-plates facing south at a
xed tilt in New York City is about 26 kWh/m2/day [22]. A typical
photoreduction electrode is composed of a semiconductor and
photocatalysts, and many of these are transition metal complexes.
Semiconductors absorb photons to make excited electrons transfer
from a valence band to the conducting band, which is then
transferred to a photocatalyst complex, which reduces CO2 to CO
and other useful organic compounds (Fig. 2). Such photoelectrocatalytic processes should be distinguished from purely
photocatalytic routes.
Kubiak and Kumar reported the photo-assisted electrochemical
reduction of CO2 to CO on Re(bipy-But)(CO)3Cl((bipy-But) = 4,40 -ditert-butyl-2,2-bipyridine)/p-type silicon with a Faradaic efciency

Please cite this article in press as: B. Hu, et al., J. CO2 Util. (2013), http://dx.doi.org/10.1016/j.jcou.2013.03.004

G Model

JCOU-6; No. of Pages 10


B. Hu et al. / Journal of CO2 Utilization xxx (2013) xxxxxx

Fig. 2. Illustration of semiconductor catalysts for photoreduction of CO2.

of 97  3%, and a short-circuit quantum efciency of 61% for light-tochemical energy conversion, and an overall efciency of about 10% for
the conversion of polychromatic light [23]. Smieja et al. further
reported that the electron transfer from the electrode to the catalyst
can be controlled by modifying the p-Si surface with phenylethyl
groups. The interaction experiments of the electrocatalyst with the
targeted catalytic substrate CO2, H2O, and CH3OH show that the
reaction with CO2 is about 25 times faster than that with H2O, and 50
times faster than that with CH3OH. Calculations based on density
functional theory (DFT) show that the nature of the binding of CO2 to
the anion forms a Re(bipy-tBu)(CO)3(CO2)K complex [24].
Kaneco has developed metal-modied p-InP photoelectrodes
for the photoelectrochemical reduction of CO2 in the LiOH/
methanol-based electrolyte. Ag, Au, Pd, and Cu deposited p-InP
photoelectrodes show higher selectivity to CO than that to H2.
Ag deposited p-InP photoelectrode show maximum current
efciency of carbon monoxide (rf = 80.4%) and Pd deposited pInP photoelectrode has the highest selectivity to CO (100%) [25].
CO is produced via one electron reduction reaction of CO2 and

CO2 is an intermediate. The Gibbs energy differences of CO(g)
splitting on many different metals change the catalytic
selectivities. Compared to electrochemical reduction at the
same metallic electrodes, onset potentials on Pb, Ag, Au, and Ni
deposited photoelectrodes are lower (about 0.250.95 V less).
The activation energy barrier in the wide-scale application
of photoreduction is the efciency of the catalyst vs. the cost of
the materials used for synthesis. If the cost of photocatalysts
and photoelectrocatalysts can be lowered and the efciencies
and lifetimes are improved, the photo-assisted electrochemical
reduction of CO2 could have possible practical applications.
3.1.2. CO production by electrocatalysis
Electricity is an easily accessible convenient energy, and can be
readily produced by a variety of renewable (wind, hydro, solar)
energy sources. The captured kinetic energy can be converted to
electricity, and then stored in the form of chemical fuels by
electrocatalysis, electrochemical reduction, electrolysis, and plasma-assisted catalytic reduction. CO2 electrocatalysis for CO

production has been conducted in either liquid electrolytes or


solid oxide electrolytes. Low operation temperature is one of the
advantages of using liquid electrolytes, but low temperature often
leads to low selectivity and low reaction rates. In aqueous solution,
the selectivity for the above carbon species is low due to the
competition of the hydrogen evolution reaction with CO2
reduction. Highly acidic electrolytes exhibit similar behavior for
CO2 reduction. Electrocatalysis cells based on solid electrolytes
(alumina, yttrium-stabilized zirconia) show high conversion rates
and selectivity to CO, but require higher operating temperatures
(around 8001000 8C) to achieve a high oxygen ion conductivity of
the electrolyte. Electrode stability based on heat expansion is also a
concern as electrodes may crack or delaminate from the electrolyte
structure.
For the aqueous CO2 reduction reaction (Eq. (3)):
CO2ad e ! CO2  aq

(3)

where CO2* is a radical anion as an intermediate. The equilibrium


potential (E vs. SCE, pH 7.0) is highly negative at 2.14 V. Ionic
liquid electrolytes such as 1-ethyl-3-methylimidazolium tetrauoroborate (EMIM-BF4) lower the potential for formation of
the CO2 intermediate, most likely by complexation via a weak
bonding between CO2 and BF4 anions. Very recently, Rosen et al.
[26] reported that CO2 can be reduced at an applied voltage of
1.5 V. This electrocatalytic system (Pt/Naon/(Ag, EMIM-BF4))
reduces CO2 to carbon monoxide (CO) at very low overpotentials
(<0.2 V). The system continuously produced CO for more than 7 h
at Faradaic efciencies greater than 96%. Another approach to
lower overpotentials is electrocatalytic reduction of CO2 by carbon
monoxide dehydrogenase (CODH), a naturally occurring enzyme
[27,28]. Shin achieved a current efciency of 100% at 0.57 V vs.
NHE in a 0.1 M phosphate buffer (pH 6.3) with no overpotential
with CODH catalysts [28].
Coordination complexes of tetraaza-macrocycles, porphyrins,
and phthalocyanines have been extensively studied for electrochemical reduction of CO2 [2932]. These catalysts have been
reviewed by Costamagna et al. [29] The key step for CO2
electroreduction is the coordination of a CO2 molecule to external
nitrogen of the phthalocyanine or naphthalocyanine ring. CO has
been produced with current efciencies up to ca. 75% with redoxinactive lithium metals [30], while the current efciencies for Co
and Fe redox metals increase up to 97.4 and 84.6%, respectively
[31]. The electrocatalytic activity of these catalysts for CO2
reduction has been enhanced by using high pressures (20 atm)
of CO2 and gas diffusion electrodes (GDE). Interestingly, high-area
metal (Fe, Ni, Cu, Pd) electrocatalysts supported on microporous
activated carbon bers (ACF) have exhibited similar effects to high
CO2 pressure, exhibiting relatively high partial current density (up
to 63 mA/cm2) for CO2 reduction on ACF/Ni compared to lower
partial current density (<10 mA/cm2) on nonactivated carbon
bers [33].
Different Pt electrocatalysts are compared in Table 1. Centi and
Perathoner developed a reverse proton exchange membrane (PEM)
fuel cell for electrocatalytic reduction of CO2 [34]. CO was
continuously produced in the gas phase in signicant amounts
(about 103 higher with respect to hydrocarbons). Nanostructured

Table 1
Comparison of Pt electrocatalysts in CO2 reduction in 0.5 M KHCO3 solution. The percentage of C10 limits the products in the trap, and signicant amounts of gas phase CO
products are not counted.
Electrocatalysts

Work conditions

Rate or I efciency

Products and selectivity

Refs.

2 nm Pt/GDM(E-TEK)
3 nm Pt/GDM
5.4 nm Pt-C/GDM
PtIr-C/GDM

Semi-half continuous, 20 8C, 20 mA, 1 h

2  102 mol/h
4.5  102 mol/h
7.4  102 mol/h
CO: 6580% H2: 1030%

10%
60%
70%
98%

[34]

80 mA/cm2, 25 8C, 2 V vs. SCE

C10, CO*
C10, CO*
C10, CO*
CO/H2 (1:2), formate < 1.5%

Please cite this article in press as: B. Hu, et al., J. CO2 Util. (2013), http://dx.doi.org/10.1016/j.jcou.2013.03.004

[35]

G Model

JCOU-6; No. of Pages 10


4

B. Hu et al. / Journal of CO2 Utilization xxx (2013) xxxxxx

carbon supported Pt nanoparticles (3 nm) catalysts have the highest


activity among three different catalysts and supports. The nanometer scale of the catalysts signicantly increased the catalytic activity.
Hydrocarbons were produced for the rst time using electrocatalytic
reduction of CO2, and the hydrocarbon production will be included
in Section 3.4. Newman and Delacourt designed a PEM electrolysiscell for simultaneous electrochemical reduction of CO2 and H2O to
make syngas CO + H2. A current density of 30 mA/cm2 has been
achieved using silver-based cathode catalysts in these electrolysis
cells [35]. Higher current density (135 mA/cm2) was achieved on
supported Au catalysts [36].
Solid oxide electrolysis cells (SOECs) have been developed for
the CO2 reduction [37]. Mogensen and Ebbesen have studied
reduction of H2O and CO2 in a Ni/YSZ electrode supported SOEC. CO
was produced via the reverse water gas shift reaction with higher
current densities (0.251 A/cm2) [38,39]. A degradation study
using impedance spectroscopy shows that ve electrode processes
contribute to the cell resistance [39]. Other SOECs using Pt paste
electrodes were much less active compared to nano-size Pt
catalysts, and no quantitative CO products have been reported
[40,41].
3.1.3. CO production by plasma
Plasmas are fast and clean methods to convert CO2 to CO and O2.
Numerous works have been performed in the plasma-assisted CO2
reduction to CO [4247]. Early work was performed in 1990s under
the international cooperation of The University of Connecticut,
Nagasaki University, and Fujitsu Laboratories Limited. The
decomposition of CO2 in fan-type ac glow discharge plasma
reactors and dielectric-barrier discharge (DBD)-plasma reactors
was investigated. CO2 decomposition was promoted by a
synergetic effect between plasma excitation in the gas phase
and catalytic actions of the metal (Au, Cu, Pt, Pd, and Rh) coated
electrode surface. CO was the main carbonaceous product with
moderate conversions at 30.5% and selectivities >80%. The energy
efciency (36%) of plasma methods need to be improved [48],
and Spencer et al. reported a theoretical energy cost analysis
method to evaluate the effectiveness of plasma systems [46]. Other
metal (Al, Cu, Ti, and Fe) electrodes for plasma activation in CH4/
CO2 were tested in Lius group [49,50], and titanium showed the
highest activity, for reasons still under investigation [51]. DBD
plasmas have been used in CH4/CO2 reforming for producing
alkanes, alkenes, oxygenates, and syngas (CO + H2) at ambient

conditions [52]. There is considerable debate on the meaning of


CO2 reforming with methane, which from an energetic and life
cycle analysis perspective appears to be a questionable approach to
limit CO2 emissions.
3.1.4. Synthesis gas production by reforming
While carbon dioxide is the primary focus of this paper, as
part of the rening process the fractions of CO2 isolated often
come with a high fraction of methane, which can be a valuable
co-reactant with carbon dioxide. Methane itself has a global
warming potential (GWP) of 30 (GWP of CO2 = 1). Methane is
abundant in shale gas, coal gas, nature gas, and landll gas. For
transportation, methane is not practical as a fuel due to its
stability and requirements for transport, though can be used in
industrial power applications. CH4/CO2 reforming to form syngas
(CO/H2) is an important process [23,24] to synthesize alternate
transportation fuels via the FischerTropsch synthesis, which
uses syngas as its feedstock. As an added benet, the
reformation of methane to fuel precursors prevents methane
from escaping into the environment and preserves carbon into
the hydrocarbon production cycle. Considerable work has been
conducted in CH4/CO2 reforming [53,54]. A few good reviews are
recommended [5557].
Power plant CO2 emission mixture has residue oxygen, which is
a poison for most CO2 reduction catalysts by promoting the watergas shift reaction. Separating the oxygen from the carbonaceous
gases adds considerable cost and complexity to CO2 conversion
efforts, thus leading to the tri-reforming process that has been
developed at Pennsylvania State University. The three-step
reaction process avoids the separation step and has the promise
of being cost-efcient for producing industrially useful synthesis
gas. The energy and environmental analysis of integrated trireforming power plants (ITRPP) has been reported by Minutillo and
Perna [58,59]. The reduction in CO2 emissions has been estimated
at 83% (15.4 vs. 93.4 kg/GJ Fuel input) and 84% (8.9 vs. 56.2 kg/GJ
Fuel input) for the ITRPP-SC (Fig. 3) and ITRPP-CC respectively. The
power plant efciency is not penalized by using the tri-reforming
process because the produced syngas can be used to generate
chemicals and/or to feed fuel cell-based power plants. Furthermore, this conventional chemical process by ammines needs high
thermal power in order to regenerate the solvent. With new
capture technology development, the lower energy requirements
for regeneration will lower the cost.

Fig. 3. Flow chart of the integrated tri-reforming power plant-steam cycle (ITRPP-SC).
Source: adapted from Ref. [58].

Please cite this article in press as: B. Hu, et al., J. CO2 Util. (2013), http://dx.doi.org/10.1016/j.jcou.2013.03.004

G Model

JCOU-6; No. of Pages 10


B. Hu et al. / Journal of CO2 Utilization xxx (2013) xxxxxx

CO2 reforming with CH4 is an example of CO2 being used as a


soft oxidant, where the dioxide is dissociated into CO and surface
oxygen, and oxygen abstracts hydrogen from methane to form
water via the water-gas shift reaction. This type of oxidative
reaction has been extended into the oxidative coupling and
oligomerization of CH4 [60,61] and dehydrogenation of alkanes
(C2H6, C3H8, etc.) [6266] and alkyl aromatics [67,68]. The
emergence, application and mechanistic pathways of these
systems have been reviewed by Park and Ansari [1]. Due to
the abundant availability, non-toxic, economic and mild oxidizing
properties of CO2, these green methods are economical and energy
efcient and they have potential industrial applications for the
development of various useful chemicals.
Shale gas contains mainly methane (60%) and light alkanes
(C2C6). With the rapid increase of shale gas production in the US
and other countries, the facile production of ethylene [63],
propene [69], butene [70], and hexene[71] from shale gas using
the method of oxidative coupling of CH4 and dehydrogenation of
alkanes is promised to lower production costs of these compounds.
3.2. CO2 conversion to HCOOH and HCHO
Formic acid (HCOOH) and formaldehyde (HCHO) are the
simplest oxygenates produced from the reduction of CO2 with
H2O (or proton solvents). Generally, they are resulting products
of CO2 reduction with protons in an electrochemical reaction,
and they are intermediates in the formation of methanol and
higher hydrocarbons. The selectivities of HCOOH, HCHO,
and CH3OH are largely dependent on the reduction methods
and catalysts used.
Yadav et al. [72] recently reported an efcient articial
photosynthetic production of formic acid from CO2 using a
graphene coupled multianthraquinone-substituted porphyrin
photocatalyst. Under 0.5 cm3/min of CO2, quantitative formic acid
products were measured with a gas chromatograph but no
quantum efciency was reported. About 46% of the total solar
light source available on earth is in the visible light range. Visible
light driven photocatalysts give an alluring prospect in the
production of fuels using solar energy, although currently most
catalysts exhibit efciencies too low for practical applications.
Graphene exhibits high electron mobility and acts as an electron
reservoir; as such a major research effort into graphenes
electrochemical and electrical properties is in effect as of this
writing. There is great potential for further accelerating the
photoreduction reaction of CO2. A signicant question here is
whether in the HCOOH synthesis by this photochemical approach
the effective mechanism involves an electrode for H2 photogeneration coupled with a catalyst to catalyze the reaction
between CO2 and H2, probably in the liquid phase. These
photocatalysts are apparently not stable.
Another photoreduction system reported by Sato et al. is
composed of a p-type semiconductor photosensitizer (N-Ta2O5)
and a Ru complex reducing catalyst in an acetonitrile/triethanolamine solution. This system has achieved a selectivity of more
than 75% for HCOOH using visible-light. The highest turnover
numbers (89) have been achieved (Fig. 4) using the N-Ta2O5
photosensitizer but the quantum efciency of 1.9% at 405 nm is
low [73].
Several different electrocatalytic reduction methods/catalysts
have been developed for continuous electrocatalytic reduction of
CO2 to oxygenate products, including HCOOH and HCHO. In
aqueous KHCO3 electrolyte, Pt and less expensive Co and Fe
nanocatalysts were respectively loaded on carbon nanotubes for
electrocatalytic reduction of CO2, with oxygenates (isopropanol,
methanol, ethanol, acetone, and acetaldehyde) isolated as the

Fig. 4. Turnover number for HCOOH formation from visible-light-induced selective


CO2 reduction using Ru complex electrocatalysts.
Source: adapted from Ref. [73].

primary products; [74] In a microuidic reactor, nanosize Pt


electrocatalysts reduced CO2 to formic acid at high Faradaic (89%)
and energetic efciencies (45%) [73].
Solid oxide electrolyzers have attracted strong interest in recent
years due to high efciency in the conversion of electrical energy
into chemical energy [37,38,7577]. Co-electrolysis of CO2 and
H2O using Ni/YSZ electrodes reported the formation of CO and H2,
attributed to the water-gas shift reaction [38]. The electrocatalysis
reactions of CO2 and H2O on Pt and ZnO nanocatalysts form HCHO
in the gas headspace, and HCHO further forms paraformaldehyde
polymers in condensed water [78]. A high CO2 conversion (8%)
was reported in the continuous gas-phase reduction. Compared to
electrocatalytic reduction in liquid electrolyte, product separation
in the gas phase reduction is not required.
Recent efforts and opportunities in the heterogeneous electrochemical conversion of carbon dioxide have been described by
Whipple and Kenis [15]. Compared to the production of HCOOH
and HCHO by photoreduction of CO2, electrocatalytic reduction of
CO2 has a much higher energy efciency (3345%); the electrocatalytic system is less complicated; also, the operation unit can be
more compact and continuously achieve high time/space yield of
the reactor. Therefore, large-scale utilization of electrocatalytic
process is technically possible. Another process relevant here is the
reduction of CO2 to HCOOH using biomass methods [139].
3.3. CO2 conversion to CH3OH
Methanol as a key commodity has become an important part of
our global economy. From building blocks for plastics, paints, and
organic solvents to clean fuels applied in fuel cells and combustion
engines, 50 million tons of methanol have been consumed in 2011.
George Olah (Nobel Prize 1994) has proposed repurposing the
hydrocarbon fossil fuel network into a methanol economy in
which methanol gathered by CO2 reduction and bioconversion is
used as a feedstock for transportation and energy storage. For
industrial production of methanol from synthesis gas (CO, CO2, and
H2), the mixture of copper, zinc oxide, and alumina has achieved
high selectivity to methanol (99.8%) at 250300 8C and 510 MPa
in the Lurgi MegaMethanol process [79] and ICI processes [80,81].
CO2 precipitates in the methanol synthesis via the reverse watergas shift reaction. Park, Lim, and their coworkers found that an
optimal CO2 fraction can maximize the methanol yield and the CO2
fraction also depends on the reaction temperature [82]. This may
indirectly show that CO2 contributes to the formation of methanol.
Not all of the equilibria have been considered in this work. The
active sites of Cu/ZnO/Al2O3 catalysts proposed by Behrens et al.
consist of Cu steps decorated with Zn atoms, all stabilized by a

Please cite this article in press as: B. Hu, et al., J. CO2 Util. (2013), http://dx.doi.org/10.1016/j.jcou.2013.03.004

G Model

JCOU-6; No. of Pages 10


B. Hu et al. / Journal of CO2 Utilization xxx (2013) xxxxxx

series of well-dened bulk defects and surface species that need to


be present jointly for the system to work [83].
Methanol can be produced directly from carbon dioxide sources
by catalytic hydrogenation and photo-assisted electrochemical
reduction. Relevant reactions for hydrogenation from CO2 are as
Eqs. (4)(6):
CO2 g 3H2 g $ CH3 OHl H2 Ol

DH298 K 49:9 kJ=mol


(4)

CO2 g H2 g $ COg H2 Og
COg 2H2 g $ CH3 OHl

DH298 K 41 kJ=mol

DH298 K 90:8 kJ=mol

(5)

Pyridinium and its substituted derivatives exhibit effective and


stable homogeneous electrocatalytic performance for the aqueous
multiple-electron/proton reduction of carbon dioxide to products
such as formic acid, formaldehyde, and methanol [95]. A p-GaP
semiconductor has reduced CO2 in pyridine to methanol with 44%
quantum efciency and near 100% selectivity at 0.5 V vs. SCE and
under 365 nm illumination [96]. Organic molecules can reduce CO2
to highly reduced species through multiple electron transfers
without the need for a metal-based multi-electron transfer. This
unique electrocatalyst shows that a single catalyst has the ability
to reduce multiple species. But the separation of methanol and
related health issues should be considered.

(6)

While the overall reaction of CO2 hydrogenation is exothermic


(DH = 49.9 kJ/mol), the rate determination step is activating CO2
in the reverse water-gas shift reaction (Eq. (5)). Transitional metal
oxides (Fe3O4, Mn3O4, and Co3O4) and Raney copper are active
RWGS catalysts at temperatures of about 300 8C. The kinetics and
mechanisms of the WGS reaction have been extensively studied on
Fe, Cu, Ni, and Au catalysts [8489]. Formate (HCOO), carbonate
(CO32), and carboxylate (HOCO) intermediates play an important role in the synthesis of hydrocarbon fuel [84]. The active sites
have been proposed by density functional theory (DFT) calculations [83,85,86]. More details about the WGS reaction can be found
[90].
Copper oxide is a catalyst for the WGS reaction. Cu/Zr/Al2O3
catalysts work for both CO and CO2 hydrogenation in the methanol
synthesis. Logically, doping metals, which function as catalysts in
the RWGS reaction, would promote CO2 hydrogenation. DFT
calculations and Kinetic Monte Carlo (KMC) simulations on metaldoped Cu(1 1 1) surfaces have shown that the overall methanol
yield increased in the sequence: Au/Cu(1 1 1) < Cu(1 1 1) < Pd/
Cu(1 1 1) < Rh/Cu(1 1 1) < Pt/Cu(1 1 1) < Ni/Cu(1 1 1) [91].
A considerable research effort has been made on CO2 activation
by visible light photocatalysts due to the natural abundance of
sunlight (Table 2). Due to the high energy requirements, this
method is often paired with electrochemical methods via
photoelectrocatalysis to push the reaction. The catalysts traditionally used are transition metal complexes, TiO2, ZnO, CdS, and
functionalized metal surfaces. While TiO2 does not have the ideal
conducting band energy for dissociating CO2 (both the CB and CO2
dissociation energies occur at 0.24 V), its anatase phase presents
a viable candidate for implementation due to its UV active band
gap (3.2 eV), its non-toxicity, and stability. A wide variety of CO2
photoreduction has been performed on the surface of TiO2 under
UV irradiation [92,93]. Mesoporous zeolite supported Ti-oxides
mainly produced methane and methanol. Addition of Pt increases
CH4 over methanol. The insight into the mechanistic aspects of CO2
photoreduction using TiO2/mesoporous materials is included in
Indrakantis review [12]. Bi2S3/CdS hetero-junction photocatalysts
exhibit a high yield of 613 mmol methanol/g catalysts under
visible light irradiation [94]. Low potentials of conduction bands of
Bi2S3 and CdS favor the formation of methanol.

3.4. CO2 conversion to long chain hydrocarbons and oxygenates


From the thermodynamic and kinetic point of view, CO2
conversion to long chain hydrocarbons and oxygenates is more
challenge than to simple CO, CH4, HCOOH, and HCHO. These
valuable liquid state fuels have higher energy density and are more
convenient to store and transport. The worlds current transportation infrastructure and vehicles are designed for consuming these
liquid fuels.
In Section 3.1, various methods for CO2/H2O conversion to
syngas CO and H2 have been introduced. Syngas can be converted
to liquid fuel via FischerTropsch synthesis (Eq. (7)).

xH2 g yCOg ! aCn H2n2 l=g bCn H2n l=g cCn H2n1 OHl
dCn1 H2n1 CHOl eCn1 H2n1 COOHl zH2 Ol

(7)

where n is a positive integer, a, b, c, d, e, and z stand for the product


fractions.
x na b c d e a c
y na b c d e
z na b c d ea  d  2e
Alkane products (CnH(2n+2)) are unbranched hydrocarbons,
suitable for diesel fuel and jet fuel. In addition, competing
reactions also produce alkenes (CnH2n), as well as oxygenated
hydrocarbons (alcohols (CnH(2n+1)OH), aldehydes Cn1H2n1CHO,
and carboxylic acids Cn1H2n1COOH). Various types of FTS
catalysts with different catalytic selectivity have been developed
for the selective production of diesel fuel, light olens, and high
alcohols. The FTS process has been used for the production of
synthetic fuels for more than 50 years. A few companies (Sasol,
PetroSA, and Shell) currently produce commercial synthetic fuels
and ne chemicals by the FTS process.
Considerable work has been reported in catalytic hydrogenation of CO2 to formic acid [97,98], methanol [99], and jet fuels [100]
using homogeneous and heterogeneous catalysts. CO2 hydrogenation using FischerTropsch catalysts has shown the advantages
over homogeneous catalysts in terms of solvent separation,

Table 2
Photoreduction and electrocatalytic reduction results of CO2 to methanol using different catalytic systems.
Catalytic-system

Working conditions

Products and selectivity

Energy eff.

Refs.

Bi2S3/CdS + CO2 + NaOH/Na2SO3/H2O


p-GaP + 10 mM pyridine + CO2

Catalysts + visible light


0.5 V vs. SCE, 365 nm

613 mmol CH3OH/g cat.


100% CH3OH

[94]
[96]

Pyridinium-KCl-H2O + CO2

Hydrogenated Pd or Pt, 50 mA cm2

TiO2/FSM16 zeolite

CO2 + H2O, 328 K

NA

[92]

Pt-TiO2/Y-zeolite

CO2 + H2O, 328 K

HCOOH
CH3OH
CH4
CH3OH
CH4, CH3OH

NA
Quant. 44%
Farad. 62%
Farad. 33%

NA

[93]

Please cite this article in press as: B. Hu, et al., J. CO2 Util. (2013), http://dx.doi.org/10.1016/j.jcou.2013.03.004

[95]

G Model

JCOU-6; No. of Pages 10


B. Hu et al. / Journal of CO2 Utilization xxx (2013) xxxxxx

Table 3
Product selectivity, olen/parafn ratio, and CO2 conversion over three different Fe FTS catalysts.
Catalysts

15%Fe/10%K/g-Al2O3
6%Fe/4%K/28%Mn
2%Fe/4%Mn/12%K/4% Ce

GHSV (h1)

3600
3360
270

Temp. (K)

673
593
563

operation cost, and catalytic activity. But different from CO


hydrogenation, CO2 is thermodynamically stable and is practically
an inert gas. The reverse water-gas shift (RWGS) reaction (Eq. (5))
is the key step for CO2 hydrogenation.
Fe [101104], Co [105,106], and Ru [107,108] catalysts
supported on silica, gamma alumina, zeolite, titania, and carbon
nanomaterials have been investigated. Addition of potassium (0.5
3%) can promote the growth of the long chain hydrocarbons
[109,110], but higher potassium contents impede the catalytic
activity, and can even kill cobalt catalysts [111]. Manganese oxide
[112] and ceria [104] are added as promoters due to the
improvement of RWGS activity [101]. Addition of Cu and Ru in
iron catalysts increases the CO2 conversion [108], and Cu sites play
the role of forming hydroxyl groups and produce higher alcohols
due to the synergetic effect among copper, potassium, and iron
catalytic sites [113]. Except catalysts design, a new concept of FTS
reactor design has been proposed by Rahimpour to increase
gasoline yield and reduce undesired products [114116].
Several efcient CO2 hydrogenation catalysts have been
recently reported for higher hydrocarbon (>C2+) synthesis
(Table 3). Without Mn promoters for the production of light
olens, the reaction temperature (where t is the absolute
temperature) in a xed bed reactor is higher as 673 K, and also
the H2/CO2 (1:3) is low [117]. With addition of MnO2 nanober
supports, the reaction temperature in a xed fed reactor decreased
to 593 K, the selectivity to C2+ hydrocarbons increased about 6%.
The interfaces of Fe/Mn supports signicantly affect the ratio of
olens/parafns. Fe supported K-OMS-2 nanocatalysts have highest selectivity to light olens (C5/C = 11.2) [118]. For the ceria
modied Fe/Mn/K catalysts in a stirred tank reactor, addition of
ceria increase 22% of CO2 conversion and 5% of increase in olen
formation due to an increase of RWGS activity compared to the
catalysts without ceria, less Fe content in catalysts still leads to
high activity for CO2 conversion due to excellent Fe dispersion
[102]. The product distributions of Mn promoted catalysts show
deviated AndersonSchultzFlory distributions due to secondary
reactions and olen adsorption and insertion [119,120].
Long chain hydrocarbons (up to C9) have been synthesized by
other CO2 conversion methods, such as electrocatalytic reduction using Pt nanoparticles on carbon-based electrodes [34]. But
the electrocatalytic reduction experiments showed that the
surface reaction-chain growth was very slow about 20 min for
the conversion of ethane to propane due to low reaction
temperature (298 K) [34]. The product distribution is different

Conv. (%)

51
45
50.4

Selectivity (%)

Refs
5
5

CO, CH4

C2C6

C /C

37.4
31.3
37.7

62.6
68.7
62.3

3.6
2.9 (11.2)
4.4

[117]
[118]
[102]

from that expected from the AndersonSchultzFlory (ASF)


distribution for FischerTropsch synthesis. This may be due to
strong readsorption of intermediates in nanostructured (porous)
carbon. Direct analyses in real time-time of ight-mass
spectroscopy (DART-TOF-MS) analysis of post-reaction catalysts
can examine the adsorbed species [118]. Syngas, light hydrocarbons, and liquid fuels have been synthesized by CO2/CH4
reforming using plasmas [52,121]. The conversions and selectivities are determined by the CH4/CO2 feed ratio, residence
time, and input power. The hydrocarbon distribution also did
not follow an ASF distribution. Carbon lms were formed
without zeolite A [121].
Copper electrodes are effective for the formation of hydrocarbons in the CO2 electrochemical reduction in the aqueous
solution (Table 4) [122124]. Ogura et al. reported that ethylene
has been produced with a high selectivity (69%) and a total current
efciency of 97% using copper (I) halide/copper electrodes
[124,125]. Copper alloy did not produce ethylene although AuCu
lowers the over potentials required [124,126]. AuCu catalysts need
to work under the same conditions for comparison. A pathway of
CO2 to C2H4 was through the formation of carboxylic acid COOH
groups and adsorbed intermediates like CH2CO, which is
supported by the formation of C1C3 carboxylic acids. Li and
Kanan reported that Cu2O layers formed on annealed Cu foil at
500 8C is more active than polycrystalline Cu electrodes [127], but
Cu2O electrodes are still less selective than copper (I) halide/copper
electrodes. 30% of HCOOH in the products are also produced,
indicating that the same pathway is followed. These experiments
suggest that Cu+ ions are the active sites for the selective ethylene
production.
3.5. CO2 as building blocks for oxygen-rich compounds and polymers
CO2 can be utilized as a monomeric building block to synthesize
various value-added oxygen-rich compounds and polymers at
mild conditions. Since polycarbonate (Eq. (8)) was synthesized by
S. Inoue and co-workers using CO2 and propylene oxide in the
presence of ZnEt2 and H2O in 1969 [128], considerable efforts have
been put into developing a copolymer synthesis based on CO2
[129131], More active and controllable catalysts [132,133] based
on Zn, Co, and Cr complexes show very high activity with a
turnover number up to 26,000, as well as excellent control for the
copolymerization of CO2 and cyclohexene (or propylene) oxide
(Eq. (9)).

(8)

(9)

Please cite this article in press as: B. Hu, et al., J. CO2 Util. (2013), http://dx.doi.org/10.1016/j.jcou.2013.03.004

G Model

JCOU-6; No. of Pages 10


8

B. Hu et al. / Journal of CO2 Utilization xxx (2013) xxxxxx

Table 4
Comparison of Cu-containing electrodes with other metal electrodes in CO2 reduction for ethylene production. Alloy 1: monel metal, Ni + Co: 65%, Cu: 33%, Fe: 2%; NA: not
available in the reference.
Electrocatalysts

Pure Cu
Cu/CuCl
Ag/CuCl
CuCl only
Alloy 1
Ag
Pt
Cu2O/Cu

Working cond.

3 M KBr, 1.8 V vs. Ag/AgCl

4 M KBr, 2 V
0.8 V, 0.5 M NaHCO3

Product selectivity
C2H4

CH4

CO

H2

51.9
69.4
64.0
42.9
0.3
1.3
49.9
5

11.4
4.0
3.8
9.9
5.2
4.9
8.0
NA

18.2
7.1
9.3
6.7
5.9
79.6
4.9
8

21.7
9.4
13.0
34.8
93.3
7.4
28.1
>50

105
97.2
98
103
108
98.4
98
NA

A high selectivity (100%) and efcient process for the synthesis


of cyclic carbonates has been developed by Yang et al. using an
electroreduction method in an ionic liquid at room temperature
(Eq. (10)) [134,135]. A current efciency as high as 90% was
achieved at a constant potential of 2.4 V. Ethylene carbonate is
used as a polar solvent with a molecular dipole moment of 4.9 D.
Propylene carbonate is a high permittivity component of electrolytes in lithium batteries.

(10)

Another breakthrough in CO2 utilization was made by Zhang


and Yu. Various propiolic acids were synthesized under mild
conditions through copper- and copper-N-heterocyclic carbene
(NHC)-catalyzed transformation of CO2 to carboxylic acid by CH
bond activation and carboxylation of terminal alkyne (Eq. (11))
[136]

CO2 conversion

Refs.

7.7
9.3
10.2
5.8
1.3
11.6
9.5
NA

[124]
[124]

[122]
[127]

CN bond formation using CO2 as C1 feedstock for the


production of various oxygenates, such as oxazolidinones,
quinazolines, carbamates, isocyanates, and polyurethanes has
been reviewed by Yang et al. [2]. These commodity chemicals have
been synthesized from green methods and have important
applications in pharmaceutical and plastic industries.
Other important examples of CO2 utilization include carboxylation of aromatics to arylcarboxylic acids with carbon dioxide
[140]. Aromatic carboxylic acids were obtained in excellent yield
by carboxylation of aromatics with a carbon dioxideAl2Cl6/Al
system at 2080 8C. The reaction mechanism was proposed to
involve superelectrophilic aluminum chloride activated carbon
dioxide reacting with aromatics via a typical electrophilic
substitution.
4. Prospective in CO2 conversion
Catalysts play a key role in the CO2 conversion to fuels and
valued added products. In the past decades, numerous methods
and catalysts have been developed to realize enhanced carbon
recovery. Current catalytic technologies for CO2 conversion have

(11)

The formation of oxazole 2-carboxylic acid in carboxylation of


oxazole (Eq. (12)) has been recently reported by Boogaerts and
Nolan using [(NHC)AuOH] complexes in THF at 45 8C [137]. The
signicantly strong Au-OH base species permit the facile
functionalization of CH bonds without the use of other
organometallic reagents.

(12)

Except for production of CC and CO bonds in the above


activation of CO2, great success has made in direct formation of
CN bonds based on CO2 activation through molecular catalysis
(Eq. (13)).

achieved high energy efciency, high reaction rates, and high value
products although these are not achieved simultaneously by a
single method. An integration of several techniques and strategies
may achieve practical production of high value chemical products
from CO2. Nanostructured, porous, and functional materials have
played and will continue to play an important role in these
catalytic conversion processes. With a better understanding of the
fundamental structurecompositionactivity relationships of
these catalytic systems, the recipes, sizes, shapes, and morphologies of the catalysts can be tuned for better catalytic performance.
Modern in situ characterization techniques and theories will
promote the rational design of cost-effective catalysts and
processes.
This review demonstrated that it is technically possible to
use CO2 as a carbon source for the synthesis of commodity
productsfrom simple CO to liquid fuels and high molecular

(13)

Please cite this article in press as: B. Hu, et al., J. CO2 Util. (2013), http://dx.doi.org/10.1016/j.jcou.2013.03.004

G Model

JCOU-6; No. of Pages 10


B. Hu et al. / Journal of CO2 Utilization xxx (2013) xxxxxx

polymers. Unfortunately, these highly endothermic CO2 conversions consume lots of energy. Renewable solar, wind, wave,
hydropower, geothermal energy, and waste heat in plants are the
rst consideration. CO2 conversion has attracted more interest due
to simultaneously reducing emissions and creating value to offset
the cost of disposal of CO2. For sustainable large-scale utilization of
CO2, the commodity products of CO2 conversion processes should
be economically viable and are in high demand. For practical
production, evaluations of energy balances and economic feasibilities of the processes are essential. The conversion processes
must also take into account the life cycle of the process to ensure
that additional CO2 is not produced beyond what is already being
removed from or going into the atmosphere.
CO2 conversion is one stage of the carbon dioxide cycle. The
compatibility of these projects and products with current
infrastructure should be considered rst and is vital for the
success of research in this area. To develop successful CO2
conversion projects, many technical and commercial barriers
need to be overcome. An interdisciplinary study should consider
overall design of product ow diagrams, economic analyses,
processes, and catalyst development and optimization. Universities and research institutions supported by government agencies
and industries mainly contribute to technology development.
Industries are commercializing these mature technologies.
Government agencies are nancially supporting this research
and development of CO2 conversion technologies, and will also
regulate and stimulate CO2 conversion technology development.
At last, CO2 utilization is a viable solution for repurposing and
storing this greenhouse gas, and international cooperation will
help mature and apply technologies to achieve the CO2 emission
control target.
The mechanisms of many of these reactions are not well
known. Various reactions are believed to have different
intermediates and mechanistic steps. One electron transfer
reactions to CO2 (to form carbon dioxide anion radical) inhibits
further conversion, while a different mechanism is necessary to
produce CO via breaking of CO bonds. Identity of the exact
steps in activation of CO2 needs to be done to improve
selectivities and yields in these reactions by modifying catalysts
based on such data. Other issues that are critical include
stability, productivity, cost, environmental friendliness of
processes and other factors.
Acknowledgements
We acknowledge the support of the U. S. Department of Energy,
Ofce of Basic Energy Sciences, Division of Chemical, Biological,
and Geological Sciences for this work under grant DE-FGO286ER13622.A000.
References
[1] M.B. Ansari, S.E. Park, Energy Environ. Sci. 5 (2012) 94199437.
[2] Z.Z. Yang, L.N. He, J. Gao, A.H. Liu, B. Yu, Energy Environ. Sci. 5 (2012) 6602
6639.
[3] C.S. Song, Role of catalysis in green chemistry for fuel processing and CO2 control,
Abstracts of Papers, 239th ACS National Meeting, San Francisco, CA, United
States, March 2125, 2010.
[4] B. Kumar, M. Llorente, J. Froehlich, T. Dang, A. Sathrum, C.P. Kubiak, Annu. Rev.
Phys. Chem. 63 (2012) 541569.
[5] B. Honisch, A. Ridgwell, D.N. Schmidt, E. Thomas, S.J. Gibbs, A. Sluijs, R. Zeebe, L.
Kump, R.C. Martindale, S.E. Greene, W. Kiessling, J. Ries, J.C. Zachos, D.L. Royer, S.
Barker, T.M. Marchitto, R. Moyer, C. Pelejero, P. Ziveri, G.L. Foster, B. Williams,
Science 335 (2012) 10581063.
[6] T.J. Crowley, R.A. Berner, Science 292 (2001) 870872.
[7] http://www.eia.gov/forecasts/ieo/ (accessed 24.12.12).
[8] http://www.350.org/en/node/26 (accessed 24.12.12).
[9] D. Damiani, J.T. Litynski, H.G. McIlvried, D.M. Vikara, R.D. Srivastava, Greenhouse
Gas Sci. Technol. 1 (2011) 111.
[10] M. Aresta, Carbon Dioxide as Chemical Feedstocks, Wiley VCH, Germany, 2010.

[11] G. Centi, S. Perathoner, R. Passalacqua, C. Ampelli, in: N.Z. Muradov, T.N.


Veziroglu (Eds.), Carbon-Neutral Fuels and Energy Carriers, 2012, pp. 291
323 (Chapter 4).
[12] V.P. Indrakanti, J.D. Kubicki, H.H. Schobert, Energy Environ. Sci. 2 (2009) 745
758.
[13] A.J. Morris, G.J. Meyer, E. Fujita, Acc. Chem. Res. 42 (2009) 19831994.
[14] J.H. Montoya, A.A. Peterson, J.K. Norskov, ChemCatChem 5 (2013) 737742.
[15] D.T. Whipple, P.J.A. Kenis, J. Phys. Chem. Lett. 1 (2010) 34513458.
[16] C.J. Liu, G.H. Xu, T.M. Wang, Fuel Process. Technol. 58 (1999) 119134.
[17] L.N. He, Z.Z. Yang, A.H. Liu, J. Gao, Preprints of Symposia American Chemical
Society, Division of Fuel Chemistry, 2011, 331332.
[18] R. Zevenhoven, S. Eloneva, S. Teir, Catal. Today 115 (2006) 7379.
[19] D.R. Lide, CRC Handbook of Chemistry and Physics Weast, Rc, 87th ed., Taylor
& Francis Group, FL, 2006.
[20] http://www.scher-tropsch.org/ (accessed 24.12.12).
[21] N.S. Lewis, D.G. Nocera, Proc. Natl. Acad. Sci. U.S.A. 103 (2006) 1572915735.
[22] http://www.nrel.gov/gis/data_solar.html (accessed 24.12.12).
[23] B. Kumar, J.M. Smieja, C.P. Kubiak, J. Phys. Chem. C 114 (2010) 1422014223.
[24] J.M. Smieja, E.E. Benson, B. Kumar, K.A. Grice, C.S. Seu, A.J.M. Miller, J.M. Mayer,
C.P. Kubiak, Proc. Natl. Acad. Sci. U.S.A. 109 (2012) 1564615650.
[25] S. Kaneco, H. Katsumata, T. Suzuki, K. Ohta, Appl. Catal. B 64 (2006) 139145.
[26] B.A. Rosen, A. Salehi-Khojin, M.R. Thorson, W. Zhu, D.T. Whipple, P.J.A. Kenis, R.I.
Masel, Science 334 (2011) 643644.
[27] A. Parkin, J. Seravalli, K.A. Vincent, S.W. Ragsdale, F.A. Armstrong, J. Am. Chem.
Soc. 129 (2007) 1032810329.
[28] W. Shin, S. Lee, J. Shin, S. Lee, Y. Kim, J. Am. Chem. Soc. 125 (2003) 1468814689.
[29] J. Costamagna, G. Ferraudi, J. Canales, J. Vargas, Coord. Chem. Rev. 148 (1996)
221248.
[30] T.V. Magdesieva, K.P. Butin, T. Yamamoto, D.A. Tryk, A. Fujishima, J. Electrochem.
Soc. 150 (2003) 608612.
[31] N. Sonoyama, M. Kirii, T. Sakata, Electrochem. Commun. 1 (1999) 213216.
[32] C.M. Lieber, N.S. Lewis, J. Am. Chem. Soc. 106 (1984) 50335034.
[33] T. Yamamoto, D.A. Tryk, K. Hashimoto, A. Fujishima, M. Okawa, J. Electrochem.
Soc. 147 (2000) 33933400.
[34] G. Wine, G. Centi, S. Perathoner, M. Gangeri, Green Chem. 9 (2007) 671678.
[35] C. Delacourt, P.L. Ridgway, J.B. Kerr, J. Newman, J. Electrochem. Soc. 155 (2008)
4249.
[36] C. Delacourt, J. Newman, J. Electrochem. Soc. 157 (2010) 19111926.
[37] M. Ni, J. Power Sources 202 (2012) 209216.
[38] S.D. Ebbesen, R. Knibbe, M. Mogensen, J. Electrochem. Soc. 159 (2012) 482489.
[39] C. Graves, S.D. Ebbesen, M. Mogensen, Solid State Ionics 192 (2011) 398403.
[40] E. Ruiz-Trejo, J.T.S. Irvine, Solid State Ionics 216 (2012) 3640.
[41] G. Tao, K.R. Sridhar, C.L. Chan, Solid State Ionics 175 (2004) 615619.
[42] S.L. Brock, T. Shimojo, M. Marquez, C. Marun, S.L. Suib, H. Matsumoto, Y. Hayashi,
J. Catal. 184 (1999) 123133.
[43] S.L. Brook, M. Marquez, S.L. Suib, Y. Hayashi, H. Matsumoto, J. Catal. 180 (1998)
225233.
[44] A. Indarto, J.W. Choi, H. Lee, H.K. Song, Environ. Eng. Sci. 23 (2006) 10331043.
[45] H. Matsumoto, S. Tanabe, K. Okitsu, Y. Hayashi, S.L. Suib, Bull. Chem. Soc. Jpn. 72
(1999) 25672571.
[46] L.F. Spencer, A.D. Gallimore, Plasma Chem. Plasma Process. 31 (2011) 7989.
[47] Z. Yoshida, H. Yosue, G. Nogami, J. Electrochem. Soc. 148 (2001) D55D59.
[48] M. Li, G.H. Xu, C.J. Liu, T. Jiang, Ranliao Huaxue Xuebao 29 (2001) 243246.
[49] C.J. Liu, L.L. Lobban, R.G. Mallinson, Stud. Surf. Sci. Catal. 119 (1998) 361366.
[50] C.J. Liu, B.Z. Xue, B. Eliasson, F. He, Y. Li, G.H. Xu, Plasma Chem. Plasma Process. 21
(2001) 301310.
[51] Y. Li, G.H. Xu, C.J. Liu, B. Eliasson, B.Z. Xue, Energy Fuels 15 (2001) 299302.
[52] B. Eliasson, C.J. Liu, U. Kogelschatz, Ind. Eng. Chem. Res. 39 (2000) 12211227.
[53] C.S. Song, P. Wei, Catal. Today 98 (2004) 463484.
[54] J.M. Wei, E. Iglesia, J. Catal. 224 (2004) 370383.
[55] J.H. Edwards, A.M. Maitra, Fuel Process. Technol. 42 (1995) 269289.
[56] D.A. Logan, Environ. Prog. 16 (1997) 237244.
[57] M.D. Ogden, G.P. Meier, K.L. Nash, J. Solut. Chem. 41 (2012) 116.
[58] M. Minutillo, A. Perna, Int. J. Hydrogen Energy 34 (2009) 40144020.
[59] M. Minutillo, A. Perna, Int. J. Hydrogen Energy 35 (2010) 70127020.
[60] A. Huang, G. Xia, J. Wang, S.L. Suib, Y. Hayashi, H. Matsumoto, J. Catal. 189 (2000)
349359.
[61] S. Tanabe, H. Matsuo, H. Matsukuma, Y. Hayashi, K. Okitsu, H. Matsumoto, S.L.
Suib, M. Hiramatsu, M. Kanamori, Sekiyu Gakkaishi 42 (1999) 383391.
[62] S. Deng, S.G. Li, H.Q. Li, Y. Zhang, Ind. Eng. Chem. Res. 48 (2009) 75617566.
[63] L. Jin, J. Reutenauer, N. Opembe, M. Lai, D.J. Martenak, S.D. Han, S.L. Suib,
ChemCatChem 1 (2009) 441444.
[64] Z.H. Shen, J. Liu, H.L. Xu, Y.H. Yue, W.M. Hua, W. Shen, Appl. Catal. A 356 (2009)
148153.
[65] X.J. Shi, S.F. Ji, K. Wang, Catal. Lett. 125 (2008) 331339.
[66] J.Q. Zhu, S. Qin, S.T. Ren, X.X. Peng, D.M. Tong, C.W. Hu, Catal. Today 148 (2009)
310315.
[67] D.R. Burri, K.M. Choi, D.S. Han, J.B. Koo, S.E. Park, Catal. Today 115 (2006)
242247.
[68] J.S. Chang, D.Y. Hong, V.P. Vislovskiy, S.E. Park, Catal. Surv. Asia 11 (2007) 5969.
[69] J. Baek, H.J. Yun, D. Yun, Y. Choi, J. Yi, ACS Catal. 2 (2012) 18931903.
[70] G. Raju, B.M. Reddy, B. Abhishek, Y.H. Mo, S.E. Park, Appl. Catal. A 423 (2012)
168175.
[71] X.B. Liu, W.Z. Li, H.Y. Xu, Y.X. Chen, React. Kinet. Catal. Lett. 81 (2004) 203209.
[72] R.K. Yadav, J.O. Baeg, G.H. Oh, N.J. Park, K.J. Kong, J. Kim, D.W. Hwang, S.K. Biswas,
J. Am. Chem. Soc. 134 (2012) 1145511461.

Please cite this article in press as: B. Hu, et al., J. CO2 Util. (2013), http://dx.doi.org/10.1016/j.jcou.2013.03.004

G Model

JCOU-6; No. of Pages 10


10

B. Hu et al. / Journal of CO2 Utilization xxx (2013) xxxxxx

[73] S. Sato, T. Morikawa, S. Saeki, T. Kajino, T. Motohiro, Angew. Chem. Int. Ed. 49
(2010) 51015105.
[74] M. Gangeri, S. Perathoner, S. Caudo, G. Centi, J. Amadou, D. Begin, C. Pham-Huu,
M.J. Ledoux, J.P. Tessonnier, D.S. Su, R. Schlogi, Catal. Today 143 (2009) 5763.
[75] Z. Zhan, W. Kobsiriphat, J.R. Wilson, M. Pillai, I. Kim, S.A. Barnett, Energy Fuels 23
(2009) 30893096.
[76] M. Ni, Int. J. Hydrogen Energy 37 (2012) 63896399.
[77] S.S. Li, Y.X. Li, Y. Gan, K. Xie, G.Y. Meng, J. Power Sources 218 (2012) 244249.
[78] B. Hu, V. Stancovski, M. Morton, S.L. Suib, Appl. Catal. A: Gen. 382 (2010)
277283.
[79] F. Pontzen, W. Liebner, V. Gronemann, M. Rothaemel, B. Ahlers, Catal. Today 171
(2011) 242250.
[80] A. Pinto, P.L. Rogerson, Chem. Eng. (New York) 84 (1977) 102108.
[81] S.G. Neophytides, A.J. Marchi, G.F. Froment, Appl. Catal. A 86 (1992) 4564.
[82] H.W. Lim, H.J. Jun, M.J. Park, H.S. Kim, J.W. Bae, K.S. Ha, H.J. Chae, K.W. Jun, Korean
J. Chem. Eng. 27 (2010) 17601767.
[83] M. Behrens, F. Studt, I. Kasatkin, S. Kuhl, M. Havecker, F. Abild-Pedersen, S.
Zander, F. Girgsdies, P. Kurr, B.L. Kniep, M. Tovar, R.W. Fischer, J.K. Norskov, R.
Schlogl, Science 336 (2012) 893897.
[84] S.D. Senanayake, D. Stacchiola, P. Liu, C.B. Mullins, J. Hrbek, J.A. Rodriguez, J. Phys.
Chem. C 113 (2009) 1953619544.
[85] A.N. Pour, M.R. Housaindokht, S.F. Tayyari, J. Zarkesh, J. Nat. Gas Chem. 19 (2010)
362368.
[86] A.A. Gokhale, J.A. Dumesic, M. Mavrikakis, J. Am. Chem. Soc. 130 (2008) 1402
1414.
[87] C.A. Cornaglia, J.F. Munera, E.A. Lombardo, Ind. Eng. Chem. Res. 50 (2011) 4381
4389.
[88] Y.Y. Chen, M. Dong, J. Wang, H. Jiao, J. Phys. Chem. C 116 (2012) 2536825375.
[89] C.S. Chen, T.W. Lai, C.C. Chen, J. Catal. 273 (2010) 1828.
[90] R.J.B. Smith, M. Loganathan, M.S. Shantha, Int. J. Chem. React. Eng. 8 (2010) 134.
[91] Y.X. Yang, M.G. White, P. Liu, J. Phys. Chem. C 116 (2012) 248256.
[92] K. Ikeue, H. Yamashita, M. Anpo, Chem. Lett. 11 (1999) 11341136.
[93] H. Yamashita, Y. Fujii, Y. Ichihashi, S.G. Zhang, K. Ikeue, D.R. Park, K. Koyano, T.
Tatsumi, M. Anpo, Catal. Today 45 (1998) 221227.
[94] X. Li, J.T. Chen, H.L. Li, J.T. Li, Y.T. Xu, Y.J. Liu, J.R. Zhou, J. Nat. Gas Chem. 20 (2011)
413417.
[95] E.B. Cole, P.S. Lakkaraju, D.M. Rampulla, A.J. Morris, E. Abelev, A.B. Bocarsly, J. Am.
Chem. Soc. 132 (2010) 1153911551.
[96] E.E. Barton, D.M. Rampulla, A.B. Bocarsly, J. Am. Chem. Soc. 130 (2008) 6342.
[97] T. Schaub, R.A. Paciello, Angew. Chem. Int. Ed. 50 (2011) 72787282.
[98] H. Takahashi, T. Kori, T. Onoki, K. Tohji, N. Yamasaki, J. Mater. Sci. 43 (2008)
24872491.
[99] S. Wesselbaum, T. vom Stein, J. Klankermayer, W. Leitner, Angew. Chem. Int. Ed.
51 (2012) 74997502.
[100] R.W. Dorner, D.R. Hardy, F.W. Williams, H.D. Willauer, 238th American Chemical
Society Meeting, Fuel Division, Washington, DC, 2009.
[101] R.W. Dorner, D.R. Hardy, F.W. Williams, H.D. Willauer, Catal. Commun. 11 (2010)
816819.
[102] R.W. Dorner, D.R. Hardy, F.W. Williams, H.D. Willauer, Catal. Commun. 15 (2011)
8892.
[103] R.A. Fiato, E. Iglesia, G.W. Rice, S.L. Soled, Iron Catalyzed CO2 Hydrogenation to
Liquid Hydrocarbons, Elsevier, 1998.
[104] F.J. Perez-Alonso, M. Ojeda, T. Herranz, S. Rojas, J.M. Gonzalez-Carballo, P.
Terreros, J.L.G. Fierro, Catal. Commun. 9 (2008) 19451948.
[105] U. Kestel, G. Frohlich, D. Borgmann, G. Wedler, Chem. Eng. Technol. 17 (1994)
390396.

[106] S.M. Kim, J.W. Bae, Y.J. Lee, K.W. Jun, Catal. Commun. 9 (2008) 22692273.
[107] N.M. Gupta, V.S. Kamble, K.R. Thampi, M. Gratzel, J. Chem. Indian Sect. A 33
(1994) 374379.
[108] S.C. Lee, J.H. Jang, B.Y. Lee, J.S. Kim, M. Kang, S.B. Lee, M.J. Choi, S.J. Choung, J. Mol.
Catal. A 210 (2004) 131141.
[109] J.S. Kim, S.B. Lee, M.C. Kang, K.W. Lee, M.J. Choi, Y. Kang, Korean J. Chem. Eng. 20
(2003) 967972.
[110] T. Riedel, G. Schaub, K.W. Jun, K.W. Lee, Ind. Eng. Chem. Res. 40 (2001) 1355
1363.
[111] G.P. Huffman, N. Shah, J.M. Zhao, F.E. Huggins, T.E. Hoost, S. Halvorsen, J.G.
Goodwin, J. Catal. 151 (1995) 1725.
[112] T. Herranz, S. Rojas, F.J. Perez-Alonso, A. Ojeda, P. Terreros, J.L.G. Fierro, Appl.
Catal. A 311 (2006) 6675.
[113] I. Boz, D. Chadwick, I.S. Metcalfe, K. Zheng, Stud. Surf. Sci. Catal. 75 (1993)
27852788.
[114] M.R. Rahimpour, M. Bayat, Ind. Eng. Chem. Res. 49 (2010) 472480.
[115] M.R. Rahimpour, A.A. Forghani, A.K. Mostafazadeh, A. Shariati, Fuel Process.
Technol. 91 (2009) 3344.
[116] M.R. Rahimpour, A. Mirvakili, K. Paymooni, Energy, vol. 36, Oxford, United
Kingdom, 2011pp. 12231235.
[117] B. Zheng, A.F. Zhang, M. Liu, F.S. Ding, C.Y. Dai, C.S. Song, X.W. Guo, Acta Phys.
Chim. Sin. 28 (2012) 19431950.
[118] B. Hu, S. Frueh, H.F. Garces, L. Zhang, M. Aindow, C. Brooks, E. Kreidler, S.L. Suib,
Appl. Catal. B-Environ. 132133 (2013) 5461.
[119] H. Schulz, M. Claeys, Appl. Catal. A 186 (1999) 7190.
[120] E.W. Kuipers, C. Scheper, J.H. Wilson, I.H. Vinkenburg, H. Oosterbeek, J. Catal. 158
(1996) 288300.
[121] T. Jiang, Y. Li, C.J. Liu, G.H. Xu, B. Eliasson, B.Z. Xue, Catal. Today 72 (2002) 229
235.
[122] K. Ogura, H. Yano, T. Tanaka, Catal. Today 98 (2004) 515521.
[123] K. Ogura, Electrochemistry 71 (2003) 676680.
[124] K. Ogura, R. Oohara, Y. Kudo, J. Electrochem. Soc. 152 (2005) 213219.
[125] H. Yano, T. Tanaka, M. Nakayama, K. Ogura, J. Electroanal. Chem. 565 (2004)
287293.
[126] Z.C. Xu, E.C. Lai, S.H. Yang, K. Hamad-Schifferli, Chem. Commun. 48 (2012)
56265628.
[127] C.W. Li, M.W. Kanan, J Am. Chem. Soc. 134 (2012) 72317234.
[128] S. Inoue, H. Koinuma, T. Tsuruta, J. Polym. Sci. Part B: Polym. Lett. 7 (1969) 287
292.
[129] G.W. Coates, C.T. Cohen, T. Chu, J. Am. Chem. Soc. 127 (2005) 1086910878.
[130] S. Fukuoka, M. Tojo, H. Hachiya, M. Aminaka, K. Hasegawa, Polym. J. 39 (2007)
91114.
[131] X.B. Lu, L. Shi, Y.M. Wang, R. Zhang, Y.J. Zhang, X.J. Peng, Z.C. Zhang, B. Li, J. Am.
Chem. Soc. 128 (2006) 16641674.
[132] S. Sujith, J.K. Min, J.E. Seong, S.J. Na, B.Y. Lee, Angew. Chem. Int. Ed. 47 (2008)
73067309.
[133] M.R. Kember, P.D. Knight, P.T.R. Reung, C.K. Williams, Angew. Chem. Int. Ed. 48
(2009) 931933.
[134] H.Z. Yang, Y.L. Gu, Y.Q. Deng, Chin. J. Org. Chem. 22 (2002) 995998.
[135] H.Z. Yang, Y.L. Gu, Y.Q. Deng, F. Shi, Chem. Commun. (2002) 274275.
[136] D.Y. Yu, Y.G. Zhang, Proc. Natl. Acad. Sci. U.S.A. 107 (2010) 2018420189.
[137] I.I.F. Boogaerts, S.P. Nolan, J. Am. Chem. Soc. 132 (2010) 88588859.
[139] G.A. Olah, A. Goeppert, G.K.S. Prakash (Eds.), Beyond Oil and Gas The Methanol
Economy, John Wiley & Sons, Inc., New York, NY, 2012.
[140] G.A. Olah, B. Toeroek, J.P. Joschek, I. Bucsi, P.M. Esteves, G. Rasul, G.K.S. Prakash, J.
Am. Chem. Soc. 124 (2002) 1137911391.

Please cite this article in press as: B. Hu, et al., J. CO2 Util. (2013), http://dx.doi.org/10.1016/j.jcou.2013.03.004

Anda mungkin juga menyukai