You are on page 1of 77

Journal Pre-proof

Current technology development for CO2 utilization into solar fuels


and chemicals: A review

Azeem Mustafa , Bachirou Guene Lougou , Yong Shuai ,


Zhijiang Wang , Heping Tan

PII: S2095-4956(20)30038-3
DOI: https://doi.org/10.1016/j.jechem.2020.01.023
Reference: JECHEM 1076

To appear in: Journal of Energy Chemistry

Received date: 5 December 2019


Revised date: 9 January 2020
Accepted date: 15 January 2020

Please cite this article as: Azeem Mustafa , Bachirou Guene Lougou , Yong Shuai , Zhijiang Wang ,
Heping Tan , Current technology development for CO2 utilization into solar fuels and chemicals: A
review, Journal of Energy Chemistry (2020), doi: https://doi.org/10.1016/j.jechem.2020.01.023

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences.
Published by Elsevier B.V. and Science Press. All rights reserved.
Review
Current technology development for CO2 utilization into solar fuels and
chemicals: A review
Azeem Mustafaa,b, Bachirou Guene Lougoua,b,c,*, Yong Shuai1a,b,*, Zhijiang Wangd, Heping
Tana,b
a
Key Laboratory of Aerospace Thermophysics of MIIT, Harbin Institute of Technology,
Harbin 150001, Heilongjiang, China
b
School of Energy Science and Engineering, Harbin Institute of Technology, Harbin 150001,
Heilongjiang, China
c
MIIT Key Laboratory of Critical Materials Technology for New Energy Conversion and
Storage, School of Chemistry and Chemical Engineering, Harbin Institute of Technology,
Harbin 150001, Heilongjiang, China
d
MIIT Key Laboratory of Critical Materials Technology for New Energy Conversion and
Storage, School of Chemistry and Chemical Engineering, Harbin Institute of Technology,
Harbin 150001, Heilongjiang, China

*
Corresponding authors.
E-mail addresses: shuaiyong@hit.edu.cn (Y. Shuai);
15bf02043@hit.edu.cn (B.G. Lougou).

Abstract
The continuous consumption of fossil fuels causes two important impediments
including emission of large concentrations of CO2 resulting in global warming and
alarming utilization of energy assets. The conversion of greenhouse gas CO2 into
solar fuels can be an expedient accomplishment for the solution of both problems, all
together. CO2 reutilization into valuable fuels and chemicals is a great challenge of
the current century. Owing to limitations in traditional approaches, there have been
developed many novel technologies such as photochemical, biochemical,
electrochemical, plasma-chemical and solar thermochemical. They are currently being
used for CO2 capture, sequestration, and utilization to transform CO2 into valuable
products such as syngas, methane, methanol, formic acid, as well as fossil fuel
consumption reduction. This review summarizes different traditional and novel
thermal technologies used in CO2 conversion with detailed information about their
working principle, types, currently adopted methods, developments, conversion rates,
products formed, catalysts and operating conditions. Moreover, a comparison of these
novel technologies in terms of distinctive key features such as conversion rate, yield,
use of earth metals, renewable energy, investment, and operating cost has been
provided in order to have a useful review for future research direction.
Keywords: CO2 conversion; Photochemical; Biochemical; Electrochemical; Plasma
chemical; Solar thermochemical processes

Azeem Mustafa is currently studying for a PhD degree from the


School of Energy Science at Harbin Institute of Technology
China. He received his B.Sc. and M.Sc. degrees in Mechanical
Engineering from University of Engineering and Technology,
Taxila in 2012 and Harbin University of Science and Technology,
China in 2018, respectively. His research interest is conversion
of carbon dioxide into solar fuels.

Bachirou Guene Lougou is Assistant professor of the School of


Energy Science and Engineering at Harbin Institute of
Technology. He received his B.S. degree in Physical-Chemistry
Science and M.S. degree in Energy Efficiency and Renewable
Energies from the University of Abomey-Calavi, Benin and
Ph.D. degree in Power Engineering and Engineering
Thermophysics in 2018 from Harbin Institute of Technology,
China. In 2018, he joined the faculty of Harbin Institute of
Technology, China. His research focuses on Carbon-neutral
energy and Hydrogen storage systems, CO2 utilization, High-
temperature solar thermochemical energy conversion, Optimum
design of redox thermochemical reactor, and Synthesis of
porous catalyst /redox oxide materials. He is also interested in
Advanced chemical energy storage materials, Electrochemical
energy conversion system, and Solar Photochemical energy
conversion system.

Dr. Yong Shuai is a full professor, Dean of School of Energy


Science and Engineering at Harbin Institute of Technology
(HIT), China. He received his B.S. degree in 2001 and Ph.D.
degree in 2006, from Harbin Institute of Technology, both in
Engineering Thermophysics. In Oct. 2004, he joined the faculty
of Harbin Institute of Technology, China. In April 2011- April
2012, he was a Visiting Scientist with the Mechanical
Engineering at Georgia Institute of Technology, Atlanta, USA.
His current research interests are in the areas of
Micro/nanoscale energy conversion and transfer, Thermal
analysis of electronic equipment and optical system, Numerical
methods for radiative heat transfer, Solar thermochemical
energy conversion, Sustainable Energy (especially solar
thermal utilization), Energy Material, Fluid Mechanics,
Computational Fluid Dynamics (CFD), and Analyzing and
testing techniques in thermal and fluid science. He is also
interested in Advanced chemical energy storage materials, and
combined Solar Photo-Thermal-Electrochemical energy storage
system.

Zhijiang Wang is a professor of the School of Chemistry and


Chemical Engineering at Harbin Institute of Technology, China.
He received his B.S., M.S. and Ph.D. degrees from Harbin
Institute of Technology. He continued his studies as a
postdoctoral fellow at the University of Oxford. His
technological interests are in the synthesis and characterization
of high-quality nanoparticles and multicomponent
nanostructures with controlled structure and interface
interactions, which could greatly enhance the potential and
broaden the application of nanomaterials in the electrocatalysis
including CO2 reduction and nitrogen reduction reactions. His
research topics presently include understanding the
mechanisms from the perspective of electronic density of state.

Heping Tan is currently a full professor and director of the


Institute of Aerospace Thermophysics of Harbin Institute of
Technology. He graduated from Harbin Institute of Technology
with a Master's degree in 1980 and obtained a doctorate degree
from Poitiers University in France in 1989. His research
interest includes Numerical heat transfer, Radiation heat
transfer, Solar thermal utilization, Analysis and suppression of
stray light in space optical system.
1 Introduction
The global energy infrastructure is transforming with potentially substantial
effects on energy trade and market because world energy utilization has high
implications for the socio-economic development of human beings [1,2]. The world
has also realized that necessary alterations will be much needed to meet the increasing
demand for energy. The current sources of fossil fuel are insufficient to fulfill the
global energy consumption, as they will get double in the next thirty years and triple
before 2100 [2]. As stated by [3], the economic break down in Europe and North
America reduced the energy consumption by 5% and 4.5%, respectively. The rates of
energy consumption were vigorous in many developing countries (>4%).
Although, the energy obtained from fossil fuel is very important because their
burning yield impressive quantities of energies/weight [4], but continuous
consumption of fossil fuels for industrial purposes is threating to the atmosphere as
large concentrations of CO2 are emitted into the environment. The energy
consumption exhibited an ascending trend from 1961 to 2010 [3]. The world is paying
huge attention to the development of renewable energy to meet the growing demand.
The main sources of renewable energy like wind, waves, sunlight, tides, biomass,
nuclear energy, and geothermal heat generate electricity with almost zero emissions of
CO2.
The abundant CO2 can be treated as an asset by employing appropriate
technologies to convert CO2 into solar fuels and chemicals. The recycling of CO2 can
reduce global warming and fossil fuel consumption [5]. There are many conversion
technologies such as photochemical, biochemical, electrochemical, plasma chemical
and solar thermochemical which can transform CO2 into valuable products.
Biochemical conversion depends on photosynthesis to transform sunlight into
biomass which can be further processed and utilized. Other novel technologies like
photochemical and solar thermochemical depend on direct solar radiation. The former
technology depends upon photons energy and the latter uses concentrated solar energy.
Plasma and electrochemical rely on electricity. The share of renewable energy in 2014
was 19.2% of the overall global energy consumption and the predicted share of
renewable energy of the world’s electricity generation was 23.7% in 2015 [6].
Currently, it is at a stage that the wide-scale utilization of renewable energy is the
main issue for better storage and convenient transportation of the electricity generated.
Though it is possible to store in batteries, it is inefficient as compared to storage in
chemicals and fuels [7]. These fuels - termed as carbon neutral or solar fuels – provide
extremely high volumetric or gravimetric storage capacity, possess high energy
densities and are compatible with existing global fuel infrastructure.
The CO2 conversion reactions can be classified into fuel production and
chemical production. The former is regarded as a highly suitable objective for the
transformation of large amounts of CO2 because it has a large market size than the
latter. Methanol is an interesting chemical and is in the middle of these two classes
because it is used as fuel and chemical in fuel cells and engines [8]. To accomplish
the conversion of CO2 into valuable products, the reactions employ co-reactants that
are actually hydrogen donors. Liquid products are better than gases for most
applications in the current existing system. The direct oxidative and indirect oxidative
routes are two approaches that can be utilized to accomplish this. The former
approach is used to produce short-chain olefins (C2H2, C3H6), oxygenated products
(CH3OH, CH2O, HCOOH) and hydrocarbons. The latter approach yields syngas
which can be further processed into almost any kind of fuel or chemical. In this
approach, the H2/CO should be controlled because it plays an important role in the
production of desired products.
Gaseous hydrogen has many serious disadvantages due to its physicochemical
characteristics as compared to liquid products [9]. Moreover, it is very costly to safely
store, dispense and transport gaseous hydrogen, while liquid products are already
compatible with the existing infrastructure. Hence, a shift from products obtained
from liquid fossil fuels (kerosene, gasoline, diesel) to a sustainable and renewable
liquid fuel is much needed. Methanol is the quite simple chemical having one carbon,
and hence can be considered as a possible solution to meet these requirements.
Nevertheless, currently, it is generated from methane for economic reasons; it can
conveniently be acquired from many carbon sources like biomass, landfill gas, biogas,
and CO2. That is why it has been suggested as essential solar fuel for the
aforementioned anthropogenic cycle of carbon [10]. While finding the most suitable
CO2 conversion approach having the potential of commercialization on a wide scale,
the efficient generation of solar fuel is very necessary to be economically compatible
with existing infrastructures. This review has discussed traditional and novel
approaches (used to produce solar fuels) to cover all the fundamental aspects along
with the latest developments followed by future perspectives for each CO2 conversion
process. A clear understanding of the matter has been presented by adding effects and
mitigation of notorious gas CO2 in the first section.
Fig.1 describes the adopted methodology and structure of this review. Initially,
effects of carbon dioxide on atmosphere and its mitigation are presented. Then a
section on carbon capture and utilization is included. Traditional thermal approaches
and their subsequent pros and cons are described in the third section of this review.
While in the fourth section of this review, photochemical, biochemical,
electrochemical, plasma chemical and solar thermochemical are discussed in terms of
their working principle, conversion rates, technological advancements, and formed
products. Finally, a comparison has been provided at the end to demonstrate the
importance of each conversion technology in certain perspectives.
Effects of CO2 and its
mitigation

Approaches for CO2


conversion
Novel thermal Conventional thermal
approaches approaches

Pure CO2
Biochemical splitting

Electrochemical Co-reactant

Photochemical Dry
Hydrogenation
reforming of
of CO2
methane
Plasma chemical

Comparison between
Solar
approaches for CO 2
thermochemical
conversion

Fig.1. Methodology of the review.


1.1 Effects of CO2 and its mitigation
The utilization of natural resources in industrial applications has
unprecedented contributions to the advancement and prosperity of human beings in
the past centuries. At present, to meet the 82% energy requirements per day, ten
billion cubic meters of natural gas, twenty-two million tonnes of coal and eighty-five
million barrels of oil are consumed [10]. These natural resources (natural gas, coal,
and oil) are carbon-rich fossil fuels and their burning leads to the emission of large
concentrations (around 30 billion) of greenhouse gas, mainly CO2 into the
environment [11,12]. The natural carbon cycle has been outpaced by these
anthropogenic emissions of CO2 and the concentrations of atmospheric CO2 have
increased from 280 ppm – before the industrial revolution – to 390 ppm in 2010 and
by the end of this century, are expected to be 570 ppm [13]. Due to the release of this
increased amount of CO2, there are detrimental effects on the environment and
climate e.g. global warming, which is having multidimensional effects on the earth’s
ecosystem and consequently affect human beings [14]. According to the report of the
Climate Resilience Handbook, numerous natural disasters – heat waves, hurricanes,
wildfires, and droughts – leading to a loss of 31 billion-dollar were recorded in 2017
[15].

Fig.2. GHG emissions by gas and economic sector-wise Production of GHG [16].
Therefore, it seems quite necessary to reduce the production of CO2 and
convert this greenhouse gas into useful chemicals for environmental protection [17].
The conversion of main greenhouse gas, CO2 into value-added products can solve the
interlinked and challenging problems of the modern age. Its conversion can not only
shift our dependence from conventional fossil fuels to newly developed fuels (Fig.2)
but also tackle the issue of global warming [18-20]. Probably, the generation of value-
added products out of CO2 can efficiently close the carbon loop.
It associates with the principle of green/sustainable chemistry if the large
anthropogenic concentrations of CO2 are reduced and converted into industrially
valuable feedstock for fuel production (Fig.3) and hence, sustained chemical industry
and energy economy can be established by developing cost-effective methods and
approaches for the mitigation of CO2 [21–23]. There are several modern CO2
conversion technologies that are currently into practice, e.g. biochemical,
photochemical, electrochemical, plasma-chemical and solar thermochemical and their
advantages and disadvantages along with their future prospects are principally
discussed in this review. However, CO2 capture, utilization, and storage (CCUS) [8,
24–30] - which are other applications of CO2 – are also briefly discussed herein.

Ethanol Formaldehyde Ethane

Products Obtained
Ethylene from CO2 Formic Acid
reduction

Carbon
Methanol Methane
Monoxide

Fig.3. Chemical products obtained from CO2 conversion.


2 Carbon capture and storage
The need to reduce emissions of CO2 from the environment is shoving
scientific community and technologists to explore and develop new strategies that
may be efficient for controlling and reducing the atmospheric level of CO2. The
capture of CO2 from flue gases of industrial processes or power plants to prevent it
from getting into the atmosphere is a possible option (Fig.4). This captured CO2 can
be utilized in different ways such as carbon source in industrial processes or disposing
of it in coal beds, oil wells, and geological cavities, etc. The CO2 capture from flue
gases released from continuous sources, that corresponds to 60% of the total released
CO2, is a mature approach. Phases of solid or liquid (membranes are also used) are
utilized to capture CO2 [31,32]. There is continuous innovation in this area and new
materials with enhanced specific characteristics are being developed like polymeric
amines, hybrid materials, metallic organic structures and functionalized silicates
which are not only energetic but are also economical [31].
CO2, O2, N2

CO2
Air Flue gases
Combustion CO2 separation
Fuel

Post-combustion capture CO2 dehydration


Compression
Separation
utilization

Gasification process
Fuel Combustion
and CO2 separation

O2 Air
CO2, O2, N2
Air Air separation

Pre-combustion capture

Combustion

O2
Recycle flue gases
Air separation
Air
Oxyfuel combustion capture

Fig.4. Capturing process for CO2 [33].


It is considered one of the most significant technologies used to decrease CO2
emissions and, in this technology, CO2 is removed directly from utility or industrial
plants and then subsequently stored in a secure medium [34]. The main CCS rationale
is to produce biofuels by reducing CO2 emissions and in connection with this, subdue
climate change [35]. In CCS technology, the CO2 storage time period should be more
than the predicted peak interval of fossil fuel usage, like if CO2 reappears into the
environment, it should arise after the estimated peak interval of fossil fuel usage.
Currently, this technology is quite promising for CO2 mitigation and is highly useful
for large amounts of CO2 [36]. Owing to the world’s big stock of low-priced fossil
fuels, there is a high need to explore new opportunities for CO2 capture from the
combustion of fossil fuel to avoid it from penetrating into the atmosphere. Large
concentrations of carbon can be obtained from several sources like coal-fired and
biomass plants. The former type of power plants is regarded as dirty because they
emit heavy concentrations of CO2, so there is an urgent need to employ CCS
technology in this type of industry. Hence, excessive amounts of CO2 will be removed
from the atmosphere and a low-pollution source of energy will be delivered to society,
emphasizing the development of efficient, economical and clean energy systems [33].
CO2 chemical recycling Synthetic
Hydrogen by water hydrocarbons
CO2 decomposition and their
separation or H2OH2 + ½ O2 products
storage Followed by methanol
synthesis
CO2 + 3H2CH3OH + H2O

CO2 capture Methanol

CO2 from
fossil fuel Fuel uses
CH3OH + 3/2 O2CO2
+ 2H2O
Atmospheric
CO2

Fig.5. Production of methanol and other chemicals via carbon conversion cycle [37].
The chemical CO2 recycling from industrial and natural resources into
different products can be obtained by utilizing CSS as a component of the reductive
hydrogenative transformation process. Fig.5 presents a carbon transformation cycle
into valuable products.

2.1.1 Working of CCS technology


Many techniques can be utilized for CO2 capture: absorption in a solid or
liquid, gas phase sequestration and hybrid processes [38, 39]. However, three main
sequestration processes can be described as [40]:
 Process 1: this process incorporates the selective CO2 uptake over a solid
surface which is reproduced by extending the temperature or decreasing the
pressure to liberate the adsorbed CO2.
 Process 2: a liquid solvent is used in this process which absorbs CO2 gas
which subsequently forms a compound. The solvent is taken to another section
and when the heat is applied to this solvent, it liberates CO2. This solvent is
further used to absorb CO2 in the next cycle.
 Process 3: a permeable membrane is used to separate CO2 from flue gases in
this method.
Pre-combustion, oxyfuel combustion, and post-combustion mechanism can be
used in capturing the process of CO2 and shown in Fig. 6 [41–43].
2.1.1.1 Pre-combustion process for CO2 capture
In this technique, CO2 is captured from the stream of syngas before the
occurrence of combustion and power generation processes. At this stage, carbon is
separated from the coal before its utilization in the stream of gas. Firstly, coal is
partially oxidized by means of gasification in the presence of air or oxygen to produce
the syngas which consists of H2+CO and methane traces. This syngas, after passing
through several phases of gas purifier is further moved through a reactor where CO is
transformed into CO2 which has high partial pressure and chemical potential paving
better ways for capture and sequestration techniques. The complete removal of CO2
synthesis gas with larger amounts of H2 can be used for combustion to generate
electricity [44,45]. More energy can be harvested from the streams of combustible gas
by using heat recovery facilities. The captured CO2 can be compressed, dried and
finally separated by physical adsorption which is a commercially adopted technique
for separation. Physical solvents are also employed in CO2 capture without
incorporating any chemical reaction but it mainly relies upon the solubility of CO2 in
the solvent. The solvent having large concentrations of CO2 can be revived by
utilizing stripping or flash desorption. Flash desorption incorporates the degassing of
a solvent having large concentrations of CO2 at high pressure through different phases
to decrease the pressure. At this stage, CO2 is separated and the solvent restores its
original shape. In stripping, after degassing, the solvent is processed with inert gas for
CO2 removal. Presently, Purisol, Selexol, Rectisol, and Fluor are commercially
available technologies that capture CO2 by physical adsorption.

2.1.1.2 Post-combustion process for CO2 capture


Many technologies are utilized to separate CO2 from flue gases released from
industrial processes. The commonly used technologies are cryogenic separation,
adsorption on membranes and solids, chemical and physical solvent. It is very costly
to establish a post-combustion system when there are large volumes of flue gases to
be processed and it seems an important limitation of this process. Other than this,
more energy is required to separate CO2 from solids and solvents after separation.
These drawbacks need serious attention before this process becomes an efficient
candidate for CO2 separation. This process needs an efficient cooling system because
CO2 captured through this process is obtained from high-temperature flue gases, low
concentrations of CO2. This process is utilized in coal-fired stations where the fuel is
combusted in the presence of air to produce streams which in turn drives large
turbines to produce electricity. The main ingredients of flue gases are CO2, O2,
impurities and moisture contents. The impurities sometimes disturb the capturing
process by damaging the sorbents [40,45].

2.1.1.3 Oxyfuel combustion


Oxyfuel combustion is another method for CO2 capture and includes the
burning of coal in the presence of O2 rather than air. The important products formed
during the process of combustion are water, CO2, and SO2, hence, CO2 separation is
not essential. CO2 capture can be carried out from flue gases by condensation of water
accompanied by compression and storage. In 1982, this method was developed for the
generation of CO2 with maximum purity for the effective removal of oil. Cryogenic
air dissociation unit provides the O2 during the early stages and appropriate
combustible conditions are maintained by mixing pure O2 with flue gases before
combustion. The high release of SO2 by flue gas may cause corrosion of the complete
setup. This process is highly advantageous in terms of technical feasibility but the
requirement of large quantities of oxygen causes high cost [46,47].

2.1.2 Challenges
Despite many advantages such as industrial CO2 utilization, carbon trading
and development of efficient CO2 storage, there are several gaps and uncertainties
concerning the utilization and development of this technology with regard to life cycle
effects, costs, storage permanence, and capacity. Furthermore, high costs of
implementation, inefficient framework and power penalties are other hindrances when
employing CCS technology.
Primarily, CCS technology is energy-intensive and costly. For instance, the
carbon capture technique itself, replacement and supply of gas sequestration materials
in energy consumption and power generation and maintenance of overall process, all
are quite expensive. It is an unfortunate fact that the CO2 capture and storage
approach needs those areas which don’t commonly exist. For instance, generally, if
the distance between the CO2 source and disposal site is about 80 km, the plant
efficiency reduces by 20%. The average storage cost mainly depends on the network
scale and facilities location. An optimized network of CO2 pipeline can minimize the
cost of infrastructure and operation as pipeline length and capacity influence the cost
of the overall system. There can’t be commercial-scale development of this
technology until the cost of implementation is reduced, which can further get public
acceptance and attract political attention.
Legislation in different countries restricts the emissions generated from
industrial applications, effectively requiring several industries to employ CCS
technology. The collaboration between different operators i.e. CCS storage and
emission source can achieve productive enforcement of CCS technology. For the
proper implementation of CCS technology, the development of a strategy concerning
the long-term utilization of this technology is necessary.

3 Conversion of CO2 – conventional thermal approaches


Before the utilization of modern CO2 conversion technologies, conventional
thermal approaches were into practice and the main difference between two was the
use of ‘fuel’. The latter used fossil fuels to provide the necessary thermal energy in
order to carry out their processes. In this section, an overview (along with their pros
and cons) of those conventional thermal approaches has been provided. Industrial-
scale conventional thermal approaches may be divided into pure CO2 decomposition
and conversion of CO2 along with a co-reactant – H2, CH4, and H2O.

3.1.1 Pure CO2 decomposition


To date, the pure splitting of CO2 is mainly ineffective due to high stability of
CO2 molecule (the Gibbs free energy of formation for C – O bond is -394 kJ/mol),
thus, requires substantial thermal energy along with a highly active catalyst and
optimal conditions for the occurrence of the chemical reaction. Moreover,
thermodynamically, both enthalpy and entropy are unfavorable for the conversion
[48-50]. The splitting reaction (R1) can be written as:
1
CO 2  g   CO  g  + O 2  g  H  293 kJ/mol (1)
2
However, there is no point of disregarding the CO2 splitting because it needs
high thermal energy to proceed, though, apparently, it seems that for such high value
of Ho that this conversion is unattainable, but there are many industrial applications
which involve endothermic reactions and can support this conversion e.g. steam
reforming of methane (R2) which is highly endothermic reaction and has worldwide
use [48-50].
CH4  g  +H2O  g   CO  g  +3H 2  g  H  206.3 kJ/mol (2)
Thermodynamically, very high values of temperature are needed for the CO2
splitting. Consider the following Fig.6 for this sake.

100 Conversion
50

(Energy effciency (%)


Energy efficiency
80 40

Conversion (%)
60 30
40 20
20 10
0 0

0 1000 2000 3000 4000 5000


Temperature (K)
Fig.6. Relation between energy efficiency and thermal conversion in terms of
temperature (reproduced from ref. [51]).
Thermal conversion and equivalent energy efficiency in terms of temperature
were discussed by Ramses et al. [51]. They theoretically calculated these values and
showed that the reaction is not very efficient at 2000 K, having reaction enthalpy 245
kJ/mol and to heat up 1 mole of CO2 from 300 to 2000 K 92 kJ/mol will be needed.
The production equilibrium of oxygen and carbon monoxide changes from <1% at
temperatures under 2000 K up to 45%–80% at 3000–3500 K.
In the past, researches have been conducted for decomposition and separation
of CO2 into various chemicals – CO and O2 – using different kinds of membranes.
Experiments were conducted by Naotsugu et al. and they utilized the yttria-stabilized
zirconia membrane to determine the CO2 splitting rate and oxygen permeation rate
using argon as sweep gas and also constituted their equations [52]. They concluded
that CO2 conversion can be improved by using the reactor system of YSZ–membrane
and they obtained only 0.5% conversion of CO2 at a temperature of 1782 K having
CO2 flow rate ranges between 4 and 250 cm3 (STP)/min. Bernard et al. studied the
production of carbon monoxide by direct decomposition of carbon dioxide [53]. Their
process consisted of a selective oxygen pumping mechanism having a semipermeable
membrane – made up of calcia – stabilized zirconia - to dissociate the carbon dioxide
equilibrium. This process produced CO in large quantities by extracting the oxygen
with the assistance of membrane, however, the overall conversion was also very low
in their study, they converted 21.5 mol% of CO2 into CO at 1954 K with equilibrium
production of 1.2%. The efficiency of the process was lowered by the material used in
the membrane which actually reduced the permeation rate of oxygen. The separation
and decomposition of CO2 by utilizing membranes made up of solid oxide was
investigated by Ren et al. [54]. In a membrane reactor, these membranes were
analyzed for CO2 decomposition into carbon monoxide and oxygen. At a temperature
of 1213 K, only 10% of CO2 was converted in the reactor with a membrane of solid
oxide (SrCo0.5FeO3) along with methane which acted as a sweep gas. Moreover, a
further 2% conversion was achieved when CO2 and helium were diluted and fed to the
lower chamber and mix up of argon and methane was sent to the upper membrane
chamber for CO2 decomposition.
The feasibility of direct splitting of CO2 into chemicals (mostly CO and O2)
with the help of semipermeable membranes for oxygen extraction has been
demonstrated in these researches. It can be concluded that the approaches presented in
these studies not only require high values of temperature but they exhibit very small
conversion rates and therefore, are not successful and efficient to be considered on the
industrial scale. Additionally, these approaches lack efficient dissociation of oxygen
and carbon monoxide which becomes an explosive blend in the presence of high
temperatures.
A concise summary of the early attempts describing several approaches for
CO2 splitting has been provided to have a complete perspective of the topic because it
has accomplished the basis for the development of novel technologies, discussed in
the next portions.

3.1.2 Co-Reactant based CO2 conversion


As discussed in the previous section that thermal decomposition of CO2
requires high energy values so the only practical approach to convert CO2 involves
the usage of a co-reactant.
From the perspective of thermodynamics, it is considerably convenient to
transform CO2 when it is coupled with a reactant having higher values of Gibbs free
energy, e.g. CH4 and H2. Basically, the inherent chemical energy of hydrogen carriers
will be utilized to support the CO2 conversion [49,50]. For CO2 conversion into
methanol and syngas, most of the conventional processes involve reaction of CO2
with either H2 or CH4 with CO2, respectively. A brief introduction to these reactions
has been included in the next sections. Since the water is an inexpensive and
ubiquitous source of hydrogen – as compared to other hydrogen sources – so the
combined reaction of water with carbon dioxide is of interest and does not have any
conventional approach and their reforming into valuable chemicals with the aid of
renewable energy will significantly impersonate the natural photosynthesis [55].
There are many other hydrogen donors which may be considered in place of CH 4, H2,
and H2O e.g. glycerol. However, in the next sections reforming of CO2 with methane
and hydrogen is presented.

3.1.2.1 Dry reforming of methane


The process of converting CO2 in the presence of CH4 is named as dry
deforming of Methane (DRM), while, in steam reforming of methane, CH4 reacts with
water instead of CO2. Fischer and Tropsch studied DRM first time back in 1928 and it
has been a challenging issue for the engineering community ever since [56]. During
world wars, there was a quest for developing alternative chemicals and fuels due to
the limited availability of fossil fuels [57–59]. After the oil crisis of the 1970s, the
scientific community started paying attention to consuming natural gas – which is
abundant and cheap – for the development of alternative fuels and chemicals in order
to reduce the dependence on fossil fuels. In the past decade, concerning
environmental protection and climate change, DRM has also been assumed as a way
to mitigate the CO2 concentration to produce value-added chemicals. The DRM
reaction (R3) is:
CO2  g  +CH 4  g   2CO  g  +2H 2  g  H  247.3 kJ/mol (3)

The dry reforming of methane is a complex process and requires high


temperature [60, 61] values – 900 to 1200 K – and some catalyst to proceed due to the
high stability of CO2 molecule and inert nature of CH4, while on the other hand, steam
reforming of methane is a pretty straightforward process due to the presence of water.
DRM, to date, is not widely utilized on the industrial scale as a major
drawback of this process is the formation and deposition of soot which results in the
early deactivation of the catalyst. The inability to efficiently convert CO 2 and
requirement of high temperature has sparked the development of new technologies –
biochemical, solar thermochemical and electrochemical, etc. for CO2 conversion. Of
course, the researches on the convenient conversion of green-house gas CO2 with the
aid of CH4 and modification of catalyst to prevent coking issue and early deactivation
are in progress [62-68].
3.1.2.2 Hydrogenation of CO2
A) Production of Methanol
The formation of methanol by selective hydrogenation of carbon dioxide is a
widely adopted process on the industrial scale across the globe and by this method,
approximately, 70 M metric tonnes are produced annually. The chemical reactions for
methanol production are given below [11]:
CO2  g  +H2  g   CO  g  +H 2O  g  H  40.9 kJ/mol (4)
CO  g  +2H 2  g   CH3OH  g  H  90.8 kJ/mol (5)
CO2  g  +3H2  g   CH3OH  g  +H 2O H  49.9 kJ/mol (6)
The production of H2O as a by-product in the methanol formation process is
the main drawback of this process because, when initiating the conversion of
methanol from synthesis gas, the one-third of the hydrogen is transformed into water.
Moreover, the formation of methanol from selective hydrogenation is not a
thermodynamically suitable reaction as compared to the production of methanol from
synthesis gas [69,70].
Regardless of the problems associated, the formation of methanol through the
reaction of CO2 and H2 is the most effective industrial-scale traditional approach for
the transformation of CO2 into valuable chemicals.
Recently, the hydrogenation of CO2 to yield C2+ chemicals has acquired a
growing interest. The C2+ species can be regarded as entry platform chemicals for
current value chains and therefore are more advantageous over C1 compounds. For
instance, C2–C4 can be utilized to enhance the calorific value of biogas and natural
gas by means of injection into distribution grids. C2–C4 (light olefins) are useful base
substances for polymer production [71]. Under atmospheric conditions, C5+ alkanes
possess high energy densities and are more suitable with the present liquid-fuel
distribution. Thus, they are the appealing forerunner of jet fuels. Corresponding to
methanol, higher alcohols are considered as clean fuels and an alternative forerunner
for small-chain olefins due to their elevated octane numbers, volatile aromatic vapors
and low NOx emissions [72]. The hydrogenation of CO2 to hydrocarbons generally
includes CO generation at first through reverse water-gas shift (RWSG) reaction
accompanied by hydrogenation of the CO intermediates. The RWSG is a key reaction
in the hydrogenation of CO2 and converts carbon dioxide into highly active carbon
monoxide by introducing hydrogen. This produced carbon monoxide can be
hydrogenated to alcohols and hydrocarbons [73].
The major hindrance in the hydrogenation of CO2 to produce C2+ chemicals is the
high coupling of C–C. Nevertheless, the encouraging discussion carried out in
different available research articles, C2+ product selectivity and time-yields has not
sufficiently been achieved. No large-scale adoption has yet accomplished. Moreover,
interactions between promotors, support and active phase are still under discussion.
More theoretical and experimental investigations are required to improve the design
of catalysts to enhance the number of active sites and their subsequent activity [74].
B) Methanation of CO2
Sabatier Reaction (R7) is an important catalytic process whereby methane is
produced by reacting H2 with CO2 in the presence of (some) catalyst and is
commercially an interesting process. Many excellent investigations about the catalytic
conversion of CO2 with H2 have been carried out [29,39,75,76].
CO2  g  +4H2  g   CH 4  g  +2H 2O H  252.9 kJ/mol (7)

Syngas and compressed natural gas are one of those many applications which
are formed by the methanation of carbon dioxide. Thermodynamically, the
methanation of carbon dioxide is quite a favorable reaction, however, this reaction has
considerable kinetic limitations because the methanation of carbon dioxide is an
eight-electron process and thus, an adequate catalyst becomes crucial to proceed the
chemical reaction [77]. This process does not have much commercial significance to
date, because 95% of hydrogen is produced through SMR and carbon capture and
utilization are costly. Methanation of carbon dioxide will be commercially more
advantageous and beneficial if carbon dioxide has been taken from streams of wastes
or inexpensive facilities and hydrogen from renewable energy sources.
C) Other products
Other than the products described in above sections, there are products such as
ethylene, formaldehyde, and ethanol which can be produced through conventional
thermal conversion of CO2. However, the process efficiency for these products is not
notable, hence, different novel technologies are used for the production of these
chemicals. For example, photoelectrochemical technology is preferably used for the
production of ethanol, photochemical reduction is employed to obtain formaldehydes
and microwave plasma is utilized for the production of ethylene. The production of
different products entirely depends on reactions conditions and the types of catalysts.
4 Novel approaches for CO2 conversion
4.1 Photochemical conversion
The irradiations of solar light are utilized for the reduction of CO2 in
photochemical transformation of CO2 and it is a sustainable and emerging approach
which has acquired huge attention as it needs low temperature, small amounts of input
energy and considerable pressure and does not have negative impact on the
atmosphere and environment [3,78]. On the other hand, other conversion approaches
– biological, electrochemical, etc. – are costly and need high values of input energy
and temperature. Because of these benefits, this approach is being employed to
overcome the current environmental issues and energy needs [79–81].
To date, photocatalytic activity and selectivity of products for CO2 reduction
are low. For the enhancement of CO2 reduction via photochemical conversion with
high selectivity and efficiency, many kinds of researches [82,83] were conducted in
the past and there have been developed a number of approaches for the improvement
of CO2 adsorption kinetics, separation and charge transfer. Few approaches are for
improvement of band-gap [84], defect control [85], development of catalyst [86],
surface sensitization [87], and modification of pore texture [88]. Engineering
processes can be utilized for efficiency enhancement of photocatalysis. This technique
can pave the method for photo-reactor designing by changing the variables such
intensity of light, selectivity, temperature and product yield [89].
Type of catalyst bed (fixed or fluidized), operation (batch, continuous) and
phase types are three important frameworks that are involved in the reactions of
photocatalytic CO2 reforming. For effective reduction of CO2 into valuable products,
the catalyst must have specificity in order to avoid the conversion of H2O into H2 and
it must reduce the energy of activation that is termed ‘potential’ in electrochemical
technology (section 4.4). The uniform photocatalytic CO2 transformation system
mainly comprises of electron relay, light absorber, the donor of electron and a
molecular catalyst [89,90] and the efficiency of such systems can be quantified by the
following terms:
 Turnover number: it corresponds to the total reductions accomplished by a
catalyst during its lifetime.
P
TN  (1)
C
Where P is products obtained after CO2 reduction and C denotes catalyst.
 Photochemical quantum yield: it is denoted by Փ and is the determination of
incident photons in a molar fraction which are converted into products from CO2
reformation.
  MP  N (2)
Where MP denotes mole products / absorbed photons and N is number of electrons
needed for conversion.
 Catalytic selectivity: it is the ratio between the product’s molar ratio after CO2
reduction and hydrogen.
P
CS  (3)
H2
Any Catalyst that produces formate and CO from CO2 transformation has the
thermodynamic capability of transforming proton into hydrogen. The value of CS can
be utilized for the quantification of efficiency of CO2 transformation with respect to
hydrogen production at particular experimental provisions. Synthetic refinements of
the catalysts give an opportunity to modify the relative reactivity with the assistance
of steric and inductive effects [90].

Photochemical
Transformation

Photosensitizer
Formic Acid

Electron Donor CO2

Methane
Catalyst

Fig.7. The photochemical CO2 reduction mechanism [89].


4.1.1 Types of photocatalytic process
CO2 reduction to useful chemicals via photochemical transformation is an
effective approach as it does not need additional energy for conversion and also does
not affect the surroundings. The process (Fig.7) of photochemical reduction of CO2
through solar energy is mainly of two types: homogeneous and heterogeneous
photocatalytic processes. The reactants and the catalyst are in the identical state of
matter in the former process while in the latter process both are in different phases.
The steps for CO2 reduction via heterogeneous photocatalysis are given below.
Step 1 It is the step of light-harvesting in which solar radiations are absorbed
by the semiconductor. For the process of photocatalytic reduction of
CO2, the energy obtained from solar radiation must be the same or
more than the bandgap energy of that semiconductor. There should be
a complete utilization of solar radiations that strike the catalyst to
avoid loss of protons and efficiency enhancement. This step is highly
dependent on the structure and morphology of the catalyst utilized for
the photocatalysis.
Step 2 The holes and protons are generated in the conductance and valance
band, respectively, as a result of the excitation of electrons in valance
band to conductance band. Tapering the gap between bands can
enhance this excitation.
Step 3 The electrons proceed to the outer area of the semiconductor and hence
restricted by the active sites of the catalyst.
Step 4 The recombining of holes and electrons takes place at this stage.
Step 5 There occurs recombination of the surface due to the recombination in
the 4th step.
Step 6 The reduction and oxidation reactions take place by using the electrons
and holes produced in the previous step.
Semi-conductors are generally used in a heterogeneous process where the
thermodynamics reactions are carried out by the solar energy for the production of
chemical fuels of high energy [91,92]. CO2 photoreduction to methanol and
formaldehyde using semi-conductors GaP, TiO2, WO, CdS and ZnO was under
research back in 1979 [92]. The holes and electrons that get to the surface of
semiconductors are restricted by the active sites of the catalyst to go through the
reduction or oxidation of the reactants and these reduction or oxidation reactions
actually transform the CO2. Once the holes and electrons are formed the feasibility of
(product) production is high [2, 93]. Similar to the process of natural photosynthesis,
solar energy can be stored in molecular bonds. Photoreactors are utilized for energy
storage whereby heat and mass transfer and reduction process take place.
Homogenous photocatalysis is not widely applicable due to the reaction’s
recombination – it should be prevented to avoid energy loss – and side reactions.
Moreover, different other process cycles may get terminated due to the degradation of
catalyst in homogenous photocatalysis.
4.1.2 Use of metal complexes for photochemical reduction of CO2
For efficient reduction of CO2, many metal complexes are being utilized due
to their several obtainable redox states and permitting the multielectron CO2 reduction.
Some examples are nickel cyclam [94], rhenium mono(bipyridyl) tricarbonyl [95],
carbine isoquinoline [96], nickel S2N2-type tetradentate [97], ruthenium
mono(bipyridyl) [98], ruthenium bis(bipyridyl) dicarbonyl [99], manganese monon
(bipyridyl) tricarbonyl [100], osmium mono(bipyridyl) dicarbonyl [101], rhodium
bis(bipyridyl) [102] etc. These metal complexes with catalysts are used in
heterogeneous systems and in these systems, the reactions are initiated by lowering
the catalyst to generate a complex metal with low valency which further reacts with
CO2. Photosensitizer provides the electron for the generation of reduced catalysts. The
most commonly used metal complexes for the photochemical conversion of CO2 are
Ruthenium and (I) and Ruthenium (II).
The main benefits of utilizing Re (I) complexes are high product selectivity
and high efficiency of reduction. Their utilization as a catalyst in photochemical
reactions only yields the CO and also restricts the generation of hydrogen. In the
1980s, these complexes were utilized as a catalyst in photochemical systems. It was
concluded in [103] that when the combination of DMF-triethanolamine solution and
Re(I) complexes yielded high selectivity and production of CO and the stability of the
photocatalysis as well CO2 reduction can be enhanced by the addition of chloride and
bromide in that solution. The CO2 reduction found to be less along with the
generation of formate complexes when chloride and bromide were not added. The
function of Triethanolamine as a reductant was significant.
The Ru (II) can be demonstrated as trans(Cl)-Ru(bpy)(CO)2Cl2, cis-
[Ru(bpy)2(CO)X]n+, [Ru (tpy)(bpy)X]n+ and as a catalyst in CO2 reduction, Ruthenium
(II) complexes are of great interest and received remarkable attention. They absorb
visible light and dissociation of CO ligand occurs when Ru complexes come in
contact with UV light, hence, photosensitizers are required for Ru (II) photocatalyst
[104]. Ru (II) catalysts particularly convert CO2 into CO and formate in the aqueous
mixture and generate hydrogen as a reduction by-product [105] while Re (I)
complexes exclusively generate CO.
Co (II) complexes also act as a catalyst in photochemistry and there two types
are [CO (macrocycle)]2+ and [CO N)n]2+. In photocatalytic reactions, they are
utilized to reduce CO2 into CO and H2. The capability of the photocatalytic process is
highly influenced by the intramolecular bonding of hydrogen, redox potential, and
macrocycle in [CO (macrocycle)]2+ complexes. In [CO N)n]2+ complexes, the
photocatalytic activity is extended by the presence of in the solution process [106].
Other than the above-mentioned metal complexes, Ir (III) and Ni (II) complexes are
also being used [107, 108].

4.1.3 Photoreactors
For the development of solar fuels by CO2 reduction, there is an increasing
interest to develop and improve the mechanism and methods for the transformation of
solar energy. The design modification of photoreactors is a new topic of interest due
to its direct effect on the efficiency of CO2 transformation.
The design of photoreactor mainly relies on light source, construction material
of photoreactor and heat exchange because the performance of photoreactor is directly
affected by these factors. A crucial step for efficiency determination of photoreactor is
as a light source. In the photocatalytic process, the range of wavelength and intensity
are measured by the light source. The photoreactor must have great capability for
photon segregation [109]. The materials utilized for the construction of photoreactors
are of many types but mainly are stainless steel, quartz, and Pyrex. Quartz is preferred
over other materials for transmission of light due to its transparency range in the
ultraviolet region. Heat exchange is also an important parameter for the designing of
photoreactor since light is the main ingredient of photoreactors so, mixing and flow
characteristics must be carefully designed for the appropriate contact between
reactants and the catalysts.

Photoreactor

Slurry Fixed bed

Membrane

1- Thin film
1- Internally illuminated 1- Fixed bed
2- Horizontal
2- Externally illuminated 2- Slurry
3- Optical fiber
4- Packed bed
5- Cylindrical
6- Monolith

Fig.8. Different types of photoreactors.


Depending upon the phase of reactants and catalysts, the photoreactors have
two types: two phase or three-phase. The phase of the catalyst may be liquid or gas in
two-phase photoreactors, on the other hand, three-phase photoreactors usually employ
reactants in gas or liquid phase with a catalyst in the solid phase. The further
classification of photoreactors is shown in the Fig.8.

4.1.4 Electron donor


The commonly used donors of the electron in the photochemical conversion of
CO2 are amines like triethanolamine and triethylamine, dihydro-1H-
benzo[d]imidazole derivatives and ascorbate, etc. In the oxidation reaction of one-
electron, a single proton is discharged by these donors to suppress the back transport
of electron from the single-electron reduction stage of the photosensitizer to single-
electron oxidation species of the donor of the electron. That discharged proton has a
powerful reductive capability to reduce catalyst, photosensitizer and the intermediates
produced. The electron donors serves as oxidative species with two-electron in the
photocatalytic system [110].

4.1.5 Catalyst
The function of a catalyst is to receive and accumulate the electron which is
emitted from the photosensitizer for the activation of CO2 by means of bonding. A
catalyst must have the ability to stifle the evolvement of hydrogen when it converges
with CO2 conversion reactions [111]. The presence of intermediates and catalysts
under the conditions of photo-irradiation is quite necessary for the excitation of the
photocatalytic process [112]. The utilization of pure catalysts, as well as their doping,
is an important area of research. Several materials in its original form don’t perform
efficiently but when they are doped with other materials, they exhibit good conversion
rates. For instance, TiO2 catalyst can’t efficiently convert CO2 and has poor selectivity
due to inadequate charge recombination and band gap (Fig.9). But when it is doped
with Ni, the production of methane increases significantly because Ni narrows down
the bandgap of TiO2 [113].
But when it is doped with Ni, the production of methane increases
significantly because Ni narrows down the bandgap of TiO2 [113]. Generally,
photocatalytic CO2 conversion primarily depends on the catalyst type (Fig.10) and
catalytic systems can be classified into three main classes: semiconductor, hybrid
systems, and complexes of transition metals. Each catalytic system has a distinct
reaction route and produces a particular product.
Methane production (u/molgcat)
16

Methane production (u/molgcat)


14 10
12 8
10
6
8 Ni-TiO2 (1.0 mol%)
6 Ni-TiO2 (0.5 mol%) 4 Ni-TiO2 (1.0 mol%)
Ni-TiO2 (0.1 mol%) Ni-TiO2 (0.5 mol%)
4
TiO2 2 Ni-TiO2 (0.1 mol%)
2 TiO2
0 0
0 10 20 30 40 50 60 70 120 130 140 150 160 170 180 190
Reaction time (min) Reaction time (min)
Methane production (u/molgcat)

Ni-TiO2 (1.0 mol%)


8 Ni-TiO2 (0.5 mol%)
Ni-TiO2 (0.1 mol%)
TiO2
6

0
0 10 20 30 40 50 60
Reaction time (min)
Fig.9. Reduction of CO2 by only TiO2 and with different amounts of nickel [114].

Photocatalysts
Surface Surface un-
supported modified
catalyst catalyst

Non-oxide Oxide Non-oxide Oxide


catalyst catalyst catalyst catalyst

Sensitized Sensitized

Flexible Flexible
substrate substrate

Metal doped Metal doped

Non-Metal Non-metal
doped doped

Fig.10. Categorization of photocatalysts for CO2 conversion.


4.1.5.1 Catalysts based on semiconductors
The use of semiconductors in photocatalytic CO2 conversion is a complex
mechanism and hard to comprehend. Broadly, there are three stages for use of
semiconductors in photocatalytic CO2 reduction, (1) production of charge-carriers (2)
dissociation and migration of pair of electron and hole and (3) redox reaction. The
working of photocatalytic CO2 conversion using semiconductor has also been
described in section 5.1.1. The use of several kinds of semiconductors in
photocatalytic CO2 conversion has been reported and titanium dioxide has been
extensively used [3]. A semiconductor will act as a catalyst if its band edge is more
negative compared to the CO2 reduction potential for a specific product [115]. While
the VB potential should be positive with regard to water oxidation potential. The
heterosturucturing approach is the most efficient pathway to increase the
semiconductor catalytic performance in photocatalytic CO2 conversion. Several
heterostructuring approaches like sensitization, co-catalyst system, charge-carrier
transfer and indirect Z-scheme acquired a lot of attention due to their ability to
enhance the performance of the semiconductor [116].

4.1.5.2 Hybrid system


This is another type of catalytic system that received huge attention for
photocatalytic CO2 conversion. In some studies, photocatalysts plus semiconductors
with different Z-scheme and sensitizer are categorized as hybrid systems. The unique
integration of semiconductors and TMC with enzymes to perform the catalytic
conversion is of practical interest because the enzyme has a strong ability of catalysis
offering high efficiency and selectivity [117]. In this catalytic system, electrons
required for CO2 reduction are provided by sensitizers which can possibly be a metal
complex or semiconductor. The hybrid system utilizing the suitable enzyme
experiences efficient production of multi-electron transport products like methane,
methanol, formaldehyde and formic acid [118]. It is observed that the nature of the
donor and the shape and size of CdS affects the catalytic activity of the hybrid system.
Many efforts have been devoted to investigating the integration of photoactive
materials and enzymes.

4.1.5.3 Catalysts based on transition metal complex (TMC)


This type of catalyst is frequently used in photocatalytic CO2 conversion
because they have long-lasting excited states and strong capability of absorbing most
part of the solar radiation [119]. Moreover, their light absorption capability and
photocatalytic activity can be changed by altering their structures. It should be noted
that photocatalytic CO2 conversion using transitional metal complex (TMC) is
strongly influenced by reactivity and structures of the (metal) complex. Mostly, three
major steps are involved in conversion when the reductive quenching route is
employed. Firstly, photocatalyst absorbs the energy of light to generate excited states
through metal-to-ligand charge transfer [120]. The metal complex works as both
photosensitizer and catalyst. Secondly, excited states will go through a reductive
quenching reaction with the donor to form a reduced complex that can be re-oxidized
to photocatalyst [121]. There are several kinds of metal complexes reported that can
be photodegraded by light radiation [122].
The poor chemical stability is the major barrier in commercial-scale use of
these metal complexes. Additionally, TMC commonly doesn’t facilitate multiple
electron-reactions because excitation (done) by a photon can only yield a single
electron [120].

4.1.5.4 Single-atom catalysts


Single-atom catalysts (SACs) having dispersed metal active centers have
recently acquired growing attention for a large class of catalytic reactions. They
exhibit extraordinary performance in CO2 conversion, and their distinctive structures
are attractive for intermediates transfer, production of required products, and
activation of CO2 molecule. Recent improvements in computational modeling,
characterization techniques, and synthetic methodologies have aided the development
of numerous single-atom catalysts. SACs mostly demonstrate remarkable
performance in a number of industrially crucial chemical reactions like photochemical,
thermochemical and electrochemical conversions. A SAC design has special
electronic and geometric characteristics which facilitate hundred percent atom usage.
The SACs reactivity is highly useful in chemo-selective hydrogenation of nitroarenes.
Recent research investigations have widely extended the space engaged by commonly
used SACs. The early examples of SACs were oxide-facilitated noble metal SACs
[123] and then single transition metal atoms grouped to N-doped carbon [124]. The
SACs preparing technologies incorporate those approaches that can utilize materials
with significant densities of surface single-atoms. The recent rapid development
indicates that such SACs will be developed which will be robust having high activity,
selectivity and stability for industrially crucial reactions [125].

4.1.6 Photosensitizer
The redox photosensitizers perform a significant role in several photocatalytic
reactions, accumulating photons for mediation of a hole or electron to catalyst from a
donor. An efficient photosensitizer should have strong optical absorption capability at
extensive wavelength ranges, mainly in the visible region and long lifetime to react
with the electron acceptor or donor. Ru (II) complexes - [Ru( N)3]2+ - are the most
extensively employed photosensitizer in different reduction-oxidation reactions e.g.
photocatalytic CO2 transformation, production of O2 or H2 from H2O [126]. The
dissociation of N in the reduction reaction is the major disadvantage of using these
complexes in the photocatalytic system. Nevertheless, the dissociated species can
work like a catalyst for carbon dioxide conversion but the process takes place in the
absence of catalyst which causes the dispersal of the product. Recently, Osmium (II)
complexes are being employed because they are powerful electron donors and have
strong absorption capability at large wavelengths close to the visible region. The other
used PS are cyclometalated Ir (III), Os (II) diamine, Re(I) carbonyl diamine, and Ru
(II) diamine [89].

4.1.7 Future perspective


It is evident from the above information that promising development has been
accomplished towards the photocatalytic CO2 transformation. However, the currently
used approaches are not sufficient to date and more efforts are needed to enhance the
efficiencies of the solar-fuel conversion. Based on the photocatalysis process, there is
a very small number of examples of chemical systems that operate efficiently. The
photon efficiency of those materials which are used in this technology and achievable
conversion rates are insufficient. Additionally, photoreactors frequently induce
operational inefficiencies in the overall system and consequently practical
applications of photoreactors are limited. In this technology, the development of
visible-light-sensitive photocatalysts which play an important role in the recycling of
large concentrations of CO2 is the immediate need. The ultimate objective is to use
solar radiation as the only input source of energy to achieve a high CO2 reduction rate.
Because visible light contains most part of the energy of the solar spectrum, it
behooves researchers to regard photocatalysts as sunlight sensitive [2, 127]. The
researches may be conducted for the better designing of co-catalysts to control and
enhance the product selectivity and uniform sensitization of co-catalysts of the
complete nanotube arrangement for improved conversion rates. The photocatalytic
CO2 reduction may become a crucial stepping stone for the production of solar fuels,
however, much development needs to be acquired before it could be adopted
efficiently on an industrial scale. It can be said that game-changing reduction rates
have not yet been attained and this approach has a long path to go to attain its full
potential to effectively compete with other approaches for solar-fuel conversion.
4.2 Biochemical conversion
The conversion of solar energy into chemicals through natural photosynthesis
for the generation of biofuels is another pathway and it is an attractive route to
transform CO2 into fuels and chemicals via biological conversion (Fig.11).
Photosynthesis can contribute significantly to CO2 conversion from several industrial
sources like chemical industries and power plants etc. since they can be employed in a
renewable and sustainable manner. Bio-fuels, food and valuable chemicals can be
generated by the direct or indirect conversion of biomass. Importantly, biomass has
many hydrogen elements and metals of alkali when being a resource of energy, the
conversion of biomass involves the same mechanism as gasification and liquefication.
If the conversion of biomass is carried out with a process of coal chemical generation,
it can accomplish synergies from the perspective of energy matching and element
supplement which will be helpful in the reduction of fossil fuel usage as well as
improvement of energy efficiency. Additionally, the emission of CO2 will be cut
down by the use of a biological system [128].
Alternative fuel consumption (%)

20

15 Biofuel
Natural gas
Hydrogen
10

2000 2010 2020 2030 2040 2050


Years

Fig.11. Projection of biofuel utilization as energy source [33].


4.2.1 Enzymatic conversion of CO2
Naturally, the CO2 fixation into organic components is not only the beginning
of biological evolution but it is also a life prerequisite. For efficient biological
evolution, six major pathways are adopted by cells and these pathways provide
encouraging reaction mechanisms for CO2 conversion. For the fixation of CO2, lyases,
oxidoreductases, and synthases contribute significantly to alter the direction and
accelerate the reaction in all six pathways. The physicochemical microenvironments
are created by cells to stabilize these complex metabolic processes (six pathways)
which also avoid the deactivation of enzymes. The vitro-based CO2 conversion
whereby an enzyme may offer an efficient route to develop CO2 capture, separation,
and utilization is of practical interest [129].
The direct use of a single enzyme (synthases, lyases or oxidoreductases) for
the accomplishment of CO2 conversion/fixation reaction for the catalytic conversion
of CO2 in vitro is quite a feasible approach for CO2 capture, separation, and utilization.
The use of Synthases in vitro application is not suitable because it involves CoA-
substrates and the obtained products are useless and invaluable in our daily lives. On
the other hand, he lyases and oxidoreductases are obtained from particular organisms
and their use help in the successful synthetization of notable materials, fuels, and
chemicals such as methane, formate, bicarbonate, and CO.

4.2.1.1 Transformation of CO2 by oxidoreductases


The transfer of electrons from one molecule to another is catalyzed by
oxidoreductase and the objective of using this enzyme is to obtain carbon-based
energy sources by reducing the oxidation state of the carbon.
CO is also undoubtedly an important fuel and chemical and considered as the
important feedstock for several synthetic mechanisms like Cativa, Fischer-Tropsch
and Monsanto processes. It also has high fuel value and can readily be transformed
into methanol. Naturally, [NiFe] CODH catalyzes the conversion between CO2 and
CO and there are centers of Ni and Fe on the active faces of [NiFe] CODH. The [3Fe–
4S] cluster connects and positions these centers so that they remain close to each other.
In the mechanism of CO2 reduction via [NiFe] CODH, reduced metal (Ni) is formed
by the transfer of an electron, accompanied by a chemical stage in which bonding
between Ni center and CO2 establishes to generate Ni-C bond. Subsequently, a
hydrogen bond is formed between histidine residue and oxygen atoms. The water loss
from Fe1 causes the production of CO2 complex whereby O1 is bound to Fe1 and
lysine residue, generates a hydrogen bond. NiIICO species are formed due to the loss
of water and the splitting of the C–O1 bond. NiIICO species quickly liberate CO and
absorbs water for the completion of the catalytic cycle. In enzymatic conversion, the
binding and catalysis of CO2 incorporates the proper location of residues and
activation by the [NiFe] centers and due to this phenomenon in vivo mechanism,
tremendous efforts have been devoted by the researchers to establish efficient vitro-
based CO2 conversion
Formate is a chemical of high importance and used for the production of
hydrogen and methanol, fuel cells and formic acid. In vivo-based CO2 fixation, the
catalysis carried out by FateDH with the NADH cofactor can convert CO2 into formate
and it is simply the transfer of hydride from the C4 atom to C atom carbon dioxide. To
facilitate this transfer, both NADH and CO2 are placed in close proximity. NAD+
along with a bipolar arrangement is left after the production of formate and is
basically the reductive acetyl-CoA route [129–131].

4.2.1.2 Transformation of CO2 by lyases


Unlike oxidation or hydrolysis, the splitting of the chemical bond is catalyzed
by lyases and then a new structure or double bond is formed. The catalysis of Michael
addition (a reverse reaction) using lyase enzyme is normally present in cells and
carboxylation of molecules along with CO2 yields Valuable products. In nature,
carbonic anhydrase (CA) is commonly found in algae, mammals, plants, and bacteria
and plays an important role in the interconversion of CO2 and bicarbonates.
Additionally, it maintains the balance of acid-base in blood and other cells of the body.
For example, [Zn] CA is used for CO2 hydration and in this mechanism, the active
sites of the metal CA comprised of Zn2+ ion are integrated with hydroxide and
histidine residues. The formation of hydroxide ligand takes place when water is
deprotonated by interacting with a weak base. This hydroxide ligand then goes
through a nucleophilic attack on the C atom to produce bicarbonates and the
molecules of bicarbonates are stabilized by H-bonding. A catalyst could be
regenerated by the displacement of bicarbonate due to water. The specific location of
CO2 in the hydrophobic section, water activation because of binding to Zn2+ and sites
of H-boding are the key factors of this method. Other than the turnover frequency of
metal CA, it is as strong as ~106 s-1. In capture and separation of CO2, carbonic
anhydrase has shown great potential due to its high efficiency of catalysis and CA has
been used, so far, for three approaches: absorption, precipitation and membrane
separation [129,132].
There exist many other kinds of lyases - other than CA - in cells that have
been used to catalyze the carboxylation of unprocessed material accompanied by CO2
in vitro. The main objective of using these lyases (decarboxylases) is to turn toxic
materials into hydrophilic and reduce their affinity for lipophilic biological
compounds. It would be interesting to note that these decarboxylases are not enzymes
included in those six major pathways and are considered as enzymes for catalysis of
the decarboxylation process. In the past, researches have been conducted on the
investigations of decarboxylases since substrate carboxylation with CO2 can
reversibly be catalyzed by these lyases [129,132].

4.2.1.3 Transformation of CO2 by microbes


Different microbes such as cyanobacteria and algae have the ability to
assimilate CO2. The different inbuilt properties of these microbes like stability,
tolerance capacity and production of metabolites are generally not suitable for CO2
transformation so their phenotype characteristics are enhanced to make them suitable
for the conversion process. The details of a few common microbes are added in the
next section.
Cyanobacteria are beneficial organisms for several industrial-scale
applications and have great potential because they are naturally convertible, require
simple nutrients and exhibit rapid cell growth. They can be used for biofuel
generation since they have natural diversity and can be grown at different locations,
even not suitable for agriculture. The transportation costs can be reduced by
producing biofuel at places close to the location of fuel consumption [133].
Cyanobacteria are found very helpful for the reduction of CO2 concentration and
enhancement of oxygen contents. Generally, they are accountable for 20%–30%
photosynthetic activity. Engineered cyanobacteria are utilized to produce ethanol, and
isoprene. In cyanobacteria, carbon utilization is mainly carried out by RuBisco,
though, its affinity is low for CO2 but has higher survival rates due to the activity of
carbonic anhydrase. The fixing CO2 efficiency of algae is low but their biomass
production is higher as compared to cyanobacteria. The RuBisco enzyme can be
genetically engineered to increase its affinity and biomass yield [89].

4.2.2 Transformation of CO2 by algae


Microalgae can be utilized to capture CO2 from different sources such as
industrial processes, atmosphere, and soluble carbonates, therefore, they can be used
for economically and technically feasible productions of biofuels (Table 1). For
biodiesel, microalgae are considered the most attractive alternative source and many
current studies and efforts are carried out on microalgae due to rich contents of oil and
high growth rate.
Table 1. Algae and their capacity of CO2 conversion.

Type of algae Conversion of CO2 Reference


μmol/(g h))
Aphanothece microscopic Nageli 136–1364 [134]
Chlorella emersonii 7–73 [135]
Chlorella Kessleri 16–155 [136]
Chlorella sp. 66–663 [137]
Chlorella sp. 109–1086 [138]
Chlorella vulgaris 591–5909 [139]
Chlorella vulgaris 335–3352 [140]
Chlorella sp. 125–1246 [141]
Euglena gracilis 36–362 [142]
Microcystis aeruginosa 49–493 [143]
Nannochloris sp 57–569 [144]
Scenedesmus obliquus 19–188 [145]
Scenedesmus sp. 39–387 [146]
Scenedesmus obliquus 52–521 [147]
Spirulina sp. 36–358 [148]
Sugars, oils and bioactive composites are also contained by algae. Other than
fuels, the improvement in technologies for efficient algal biodiesel generation is also
suitable for biogas, bioethanol, biohydrogen and biomass-to-liquid technologies
employing fast-growing microalgae. Additionally, other useful products e.g., animal
feed, biopolymers, and proteins can be generated during this process. Due to the small
harvesting cycle, multiple harvests with more yields can be produced from algae.
Moreover, the cultivation of algal can be carried out on non-arable or marginal land
with the use of wastewater [149].

4.2.3 Future perspective


Biochemical processing of CO2 has been an emerging technology to mitigate
global warming and produce chemicals and fuels. Based on the natural CO2 metabolic
mechanism, two types of enzymes are being used successfully for CO2 transformation
into value-added products. There have been reported several methods for efficiency
enhancement of enzyme-based catalytic processes. Although a lot of work has been
done, significant technical advances and investigations are required to employ
enzymes for CO2 conversion on large-scale. The enzymes which are being used
nowadays can’t be successfully used on a commercial scale because they are
expensive and possess high sensitivities with respect to the environment. Some
developed techniques and their scaling up to commercial level is highly needed.
Enzymatic transformation of CO2 involves reactions that are cofactor dependent, the
price of these cofactors is high plus their availability is less which restricts their
utilization on large-scale applications. Hence, more researches should be devoted to
not only excavate new low input energy consuming and inexpensive techniques for
the effective reuse and regeneration of co-factor but enzymatic reactions pathways
without using cofactors may also be developed.
Nevertheless, biofuel resource has huge inherent potential, the large-scale
production of algae for the generation of biofuels is still in progress. The
identification and development of economical methods for the growth of oil-rich
algae and reduction of cost using integrated bio-refineries to generate biogas,
biodiesel, electrical power and animal feed can significantly improve the potential of
this technology. Moreover, the use of nutrients from industrial wastewater and
municipal sewage to minimize the consumption of fertilizers for algae growth,
determination of area-efficient methods for CO2 capture and development of energy-
efficient and inexpensive harvesting techniques can play an extremely important role
to make this technology commercially viable and meet the energy demands.

4.3 Electrochemical conversion


The electrochemical conversion of CO2 has gained great attention due to its
various advantages. This process is controlled through reaction temperature and
electrode potentials and the overall consumption can be reduced because electrolytes
can be completely recycled. Additionally, this conversion process is modular,
compact, on-demand and can easily be scaled up for different applications and
electricity required to drive the system can be acquired without producing new CO2
[150].
Carbon Monoxide
(CO)

(HCOOH) or formate
(HCOO-)

Electrochemical reduction of CO2


(H2C2O4) or (C2O42-)

Methanol (CH3OH)

Formaldehyde
(CH2O)

Methane (CH4)

Ethanol
(CH3CH2OH)

Ethylene
(CH2CH2)

Fig.12. Products generated by electrochemical transformation of CO2.


However, there are challenges, as well. For instance, slow kinetics of CO2
reduction and small energy efficiencies even at higher electrode potential. The
electrochemical reduction process also consumes high energies. The electrocatalysts
exhibit poor performance in terms of stability and catalytic activity. The reaction
mechanisms for electro-reduction of CO2 are very complicated even if catalysts are
used and generally, catalysts being used are not much active. Depending upon the
reaction medium and material of the catalysts, the electrochemical transformation of
CO2 yields different kinds of products (Fig.12) via two, four, six and eight electron
reduction routes in electrode configurations. The products that can be produced via
electrochemical transformation of CO2 are influenced by the materials of electrode
and catalysts, pressure and concentration of CO2, reaction medium and temperature,
pH, buffer strength and electrolyte solution [17].

Types of
electrolysis Cell

(SPCEC)
(SOEC) Alkaline
Solid
Solid oxide electrolysis
proton conducting
electrolysis cell cell
electrolysis cell

Fig. 13. Types of the electrolysis cell.


Fig.14. Working principles of SOEC and SPCEC [51].
The design of the electrochemical cell is an essential aspect that has an impact
on the process of CO2 reduction, ascertain Faradaic efficiency, stability, and current
density. Electrolysis cells are mainly of three types (Figs. 13 and 14); SOEC and
SPCEC are used for the combined/pure and combined transformation of H2O and CO2,
respectively. While water splitting is carried out in the third type of cell. Since the
first report on CO2 reduction in 1980, different cell designs have been proposed for
efficient reduction. Currently, reactor vessels can be generally categorized into H-type
cells, microfluid cells, SOEC, DEMS cells, and polymer electrolyte membrane flow
cell.

4.3.1 Catalysts
As discussed in section 5.1.1, the function of a catalyst is to proceed with the
selective production of the desired product and to connect and activate CO2 to
minimize the high overpotentials. The choice of catalyst is highly important for the
selective production of different products on a commercial scale. In CO2 electro-
reduction, elements of transition metal and related compounds are commonly used as
electrocatalyst due to the presence of active electrons and vacant orbits which
energetically assist the bonding between CO2 and metal. Four distinct streams of
metal catalysts are significantly used for the electrochemical transformation of CO2.
At ambient temperature and pressure, copper catalysts have a unique
capability of CO2 reduction into hydrocarbons. Modification of Cu surfaces has
exhibited an increase in selectivity for hydrocarbon production and reduction of
overpotential. Cu electrode coated with Cu nanoparticles has also shown impressive
selectivity for hydrocarbon production [151]. Formic acid with higher Faradaic
efficiency can be produced by using metals for hydrogen production e.g. Pb, Hg.
MOF catalysts, metal oxides, and alloys significantly contribute to the production of
HCOOH. Methanol is a valuable chemical due to its many applications. It can be
produced by reduction of CO2 through different processes such as heterogenous (more
successful) and electrochemical methods (low faradic efficiencies) [152]. Carbon
monoxide can be generated efficiently on several metal electrodes and syngas can be
formed when it combines with H2 through the Fischer-Tropsch process. The negative
potential for CO production is higher than H2 so most researches focus on enhanced
CO production using more developed systems. The use of nanoparticles is also
efficient for CO productions [153].

4.3.2 Components of an electrochemical reduction system


4.3.2.1 Electrolyte
Although electrolytes strongly affect every electrochemical system, few studies to
date have paid attention to the effects of its composition [153]. The heterogeneous
process for the electrochemical transformation of CO2 generally uses aqueous
electrolytes which are composed of anions (bicarbonate, hydroxide, and halide) and
cations (Sodium+, Potassium+). These salts are frequently utilized because they
possess high conductivities in water, and proton needed for proton transfer in the
electrochemical process is offered by the water present in the electrolytes [153]. The
selection of electrolyte strongly affects the efficiency of CO2 reduction, current
density, and product selectivity. For the same electrolyte, the selection of anion will
affect the pH of the solution at the Cu electrode which in turn, will strongly influence
the amounts of products generated. Similarly, the selection of cation for bicarbonates
(electrolyte) will influence the production distribution on the Cu electrode. The
buffering capacity of anion and size of cation also have impact on pH and
overpotential (the difference in voltages of electrodes between thermodynamic and
actual world values to proceed a reaction), respectively [154, 155].

4.3.2.2 Electrode
Electrodes also play an important role in each heterogeneous electrochemical
reduction of CO2. An electrode is composed of the catalyst layer and baking substrate
of performing multiple functions like reactant gas delivery to the catalysts, transfer of
product produced from catalyst to electrolyte membrane. Enhancing electrode and
reactor performance requires optimization of such processes which depend on the
intricate structure of the electrode. The use of electrodes made up of planar metals
(Cu, AU, Ni, Sn, Ag, Ti, etc.) is restricted due to lower CO2 concentrations and
surface area because of insufficient CO2 solubility in aqueous electrolytes [156].

4.3.2.3 Gas adsorbent material


Conjugated microporous Polymer (CMP) is widely used for adsorption of gas
and capacity applications owing to their micro-porosity associated with the large
surface zone. The application of the above-mentioned polymers in CO2 reduction is
another area of advancement and few references are available about their employment
as terminals. The feasible congregate of the multipurpose cathode is to provoke the
importance to generate efficiency, selectivity, and effectiveness of the electrocatalytic
CO2 transformation. Desired morphology, sufficient surface territory, prevalent
usefulness, and sane pore-estimate dispersion are the important variables for the
transportation of electron and mass and energy of substance response. Metal-organic
frameworks in comparison with other permeable materials exhibit exceptional sizes of
the tunable pore, multifunctionality, and permeable structure which strengthen them
to compete with other permeable materials in different applications. Additionally,
their design is helpful for the definite and direct legitimization of the associated
structure-property relationship which is considered an important aspect of materials
enhancement. The exceptional advancement of these MOFs has gained attention from
many areas and they are animating the progress of wave of frequent exuberance
applications. The extraordinary limit of gas storing of MOFs at a different range of
temperatures has resulted in their rapid development, making enormous adsorbents
highly useful in energy components, power offices, vehicles, and gas tank [89, 157,
158].

4.3.3 CO2 conversion by microbial electrolysis


The products that are formed as a result of electrochemical CO2
transformation are largely C1 species such as CO and HCOOH. Although several
electrocatalysts can be utilized to achieve the coupling of C-C which is a difficult job
to accomplish, microbial electrolysis is preferably being used for this sake [159]. It is
a novel approach for transforming renewable sources of energy such as solar and
wind energy into useful chemicals and fuels from CO2. In this approach, microbes are
used to convert CO2 into organic species with electrons (obtained from cathode). For
the development of microbial electrolysis, microbes and materials of the electrode are
considered important factors. Mixed and pure cultures are two types, utilized for CO2
transformation in microbial electrolysis. The mixed consortium is mainly separated
from sludge or wastewater, etc. and yields acetate which is a base for the formation of
hydrogen, formate, and methane. In the case of a mixed consortium, there is no issue
of cell apoptosis and the need for sterile conditions [160]. In microbial electrolysis,
cathodes are made up of carbon-based materials and cathode improvements like
surface chemistry and area are in progress to optimize microbial electrolysis. This
approach has good potential for CO2 transformation and researches is being
conducted for its development. The development of efficient and inexpensive
cathodes can enhance the effectivity of this system.

4.3.4 CO2 conversion by photoelectrochemical cell


An emerging technology for CO2 conversion is photoelectrochemical cell (PEC). It
combines electrocatalysis and photocatalysis approaches because there are distinct
areas of interest of photoelectric synergistic reduction of CO2 with water. Firstly, the
usage of solar energy can primarily lower the connected voltage, in this way
decreasing the power consumption, and other than this, the load of external
predisposition voltage can steer the detachment of holes and electrons
(photogenerated), which is a basic advancement deciding the productivity of
photocatalysis. Most of the time, the methods frequently embraced for the H2O
portion have been misused for the conversion of CO2. The CO2 initiations are
convoluted as compared to water in many cases. The CO2 reduction using
photogenerated electrons can be an aggressive approach with H2O reduction, in this
route making necessary for these two technologies vary sufficiently. The CO2
reduction yields more entangled products than the reduction of H2O which only
produces H2. The route of the photoelectrochemical transformation of CO2 to
attractive chemicals emphasizes CO2 conversion from early investigations to ongoing
revelations, underscoring the refinements on the latest pathways in major
comprehension of the structural effects of the cocatalysts and semiconductors on
photoelectrochemical CO2 reduction and several products arrangements [161, 162].

4.3.5 Impact of different factors on product selectivity


Product selectivity can be mainly determined by the reduction route and
process kinematic which results in the generation of different chemicals. The
electrons are obtained by CO2 either from cathode or medium in the course of early
phases of electrochemical reduction of CO2. As a result, an intermediate is formed
which is absorbed by the cathode for the production of desired products. The nature
and type of reaction route influence the electrode potential, catalyst, electrolyte
solution, pH and concentration of CO2 [163].
The bonding accomplished between carbon and adsorbed anion, increased the
electron transfer from anion to carbon dioxide, hence, initiating the conversion of CO2
when electrode composed of copper mesh was placed in the solution. The stronger
halide anion adsorption to the electrode help achieve the highest CO2 conversion. The
use of cations like cesium and rubidium also enhances the electrochemical reduction
of CO2 to CO. The use of different metals (for electrodes) and their corresponding end
products are shown in the following Tables 2–5 [164].
Table 2 Metals and their end products
Metals End product
Ru CH3OH
Ag, Au CO
Zn CO + HCOOH
Cu CH4, H2, C2H2, CO, HCOOH subjected to the
conditions
Cd, ln, Pb, Sn HCOOH + CO

It is found highly suitable to employ an electrode made up of single metal for


the electrochemical reduction of CO2 [165]. In Table 3, the use of particular metals
and end products is given. The different other electrocatalyst materials are also used
for the production of specific products with maximum current efficiencies. The poor
water solubility (at STP) restricts the utilization of CO2 in an aqueous solution.
Table 3. Catalysis of Electrodes by different metals with aqueous and nonaqueous electrolytes.
Metal End result
Aqueous Electrolyte Nonaqueous
electrolyte
sp metals (ln, Sn, Cd, Zn, Cu, Ag, Au) CO
sp metals (Hg, ln, Sn, Pb) Formate Oxalate
d metals (Pt, Pd) CO also HCOOH, HCHO, CO including
CH3OH and HC Ni
d metals (Fe, Cr, Mo, Ti, Nb) Oxalate and
CO
d metals (Ru) HC, formaldehyde, alcholols
Table 4. Electrodes working with aqueous and non-aqueous electrolytes.
Metal End result
Aqueous electrolyte
Zn, Au, Ag CO
Cu Hydrocarbons, aldehydes alcohols
ln, Sn, Hg, Pb Formate
Al, Ga, Group VIII (except Pd) Weak performance
Non-aqueous electrolyte
Cu, Ag, Au, Sn, ln, Zn CO and CO32-
Ni, Pd, Pt CO only
Pb, Tl, Hg Oxalate
Al, Ga, Group VIII (except Ni, Pd, Pt) CO and Oxalic

Only small CO2 quantities are available on the surface of the electrode to
procced the reaction. However, this problem can be overcome by using pressurized
CO2. Metals like ln or Pb generate more CO while Pb and Cu increase the production
of formic acid at pH 4.0 [164].
Table 5. A summary of experimental conditions, catalytic materials and main products.
Catalyst Type of Composition of Current density Pressure/ Main Reference
electrolyte electrolyte (mA temperature product
cm−2)
Ag Solution + 18% mole − Ambient/ RT CO (96%) [166]
CO2 gas EMIM
BF4 in water
Pd, Ag and Zn Solution + 10 mM KHCO3 ~ 200 Ambient/ RT CO [167]
CO2 gas (> 90%)
Ag foil CO2 0.1 M KHCO3 4 (-1.02 V Ambient/ RT CO [168]
Saturated vs RHE) (90%)
solution
An-Cu CO2 0.1 M KHCO3 20 Ambient/ RT C2H4 [169]
Saturated (-1.08 V vs. (38.1%)
solution RHE)
Cu/Ag CO2 0.05 M − Ambient/ RT C1 and [170]
Saturated CS2CO3 C2/CO
solution
CoPc CO2 gas 1 M KOH 150 Ambient/ RT CO (95%) [171]
Solution +
CO2 gas
Graphite/Cu/ C Solution + 7 M KOH ∼100 Ambient/ RT C2H4 [172]
NPs/PTFE CO2 gas (-0.55 V vs. (70%)
RHE)
Cu2O/ZnO Solution + 0.5 M KHCO3 − Ambient/ RT CH3OH [173]
CO2 gas (25.6%)
ln/Pb CO2 1 M NaHCO3 40 Ambient/RT HCOOH [174]
Saturated (80%)
solution
Cu NPs CO2 0.1–0.5 (-1.8 V vs. 1 – 9 atm /RT C2H4 + [175]
Saturated KHCO3 Ag/AgCl). CH4
solution (45%)
RuPd/Sn CO2 gas 1 M KOH + ∼100 Ambient/ RT HCOOH [176]
Solution + 0.5 M KCL (-0.55 V vs. (89%)
CO2 gas RHE)

CuNp Solution + 1 M KOH ~200 Ambient/ RT C2H4 [177]


CO2 gas (-0.7 V vs. (46%)
RHE)

Cu nanocube CO2 0.25 M KHCO3 68 Ambient/ RT C2 (over [178]


Saturated (-0.96 V vs. 60%)
solution RHE)

NS-C CO2 0.1 M KHCO3 2.86 Ambient/ RT CO (92%) [179]


Saturated (-0.49 V vs.
solution RHE)

Ni/YSZ H2O - CO YSZ − 750 °C CO [180]


mixtures or electrolyte
CO2-H2 tubes

4.3.6 Future perspectives


To date, enormous rewards and potential have been accomplished by
electrochemical CO2 reduction. But this approach is not mature enough for the
production of low carbon fuels on an industrial scale due to various technological
challenges like high overpotential, the generation of entangled products, material
instability of electrodes, high consumption of energy, small Faradic efficiencies,
kinetic barriers and low CO2 solubility in aqueous solution. Generally, the biggest
challenge is the poor performance of electrocatalysts because of insufficient stability
of catalysts, poor catalyst activity, and product selectivity. There is no electrocatalyst
available to date for CO2 conversion on a large-scale. Hence, stable catalysts having
high selectivity and catalytic activity should be developed on priority bases. New
approaches and methods for the activation of CO2 molecule at small values of
overpotentials are needed to effectively transform CO2 into valuable chemicals and
fuels. The novel electrodes should be developed which enable electrolysis at suitable
current densities. Novel electrodes allowing electrolysis at current densities near to
water electrolyzers (which are commercially available) should be developed, for
which electrodes composed of solid oxide become suitable options. Moreover, it is
needed to better understand the mechanistic role of metal oxides in order to generate
the possibility of electrode design with different compositions. It can be concluded
that efforts should be devoted to the development of durable catalysts as well as
optimization of system design should be carried out. Several research directions in the
future may include improvement in stability and catalytic activity by exploring new
electrocatalysts, optimization of reactors, electrodes and system design for large-scale
applications.
4.4 Plasma technology
Many efforts have been devoted to the development of energy-efficient
methods and technologies to convert CO2 in value-added products. Plasma technology
is one of these emerging technologies which are gaining a lot of interest. Plasma is
actually a (partially) ionized gas and consists of electrons and numerous neutral
species such as different kinds of ions, radicals, molecules, and excited species. The
plasma is an extensively reactive chemical due to the interaction of the above-
mentioned species with each other [181, 182]. The high potential of this technology
for the conversion of CO2 has a huge potential due to the existence of energetic
electrons in a plasma. Plasma is generated when gas is broken down by employing
electric power to it, i.e., the generation of positive ions and electron. Electrons have
low weights as compared to other plasma species and most energy obtained by
electrons is from the electric field. They have higher energies and don’t consume their
energy so expeditiously in collision with other species. Thus, these high energy
electrons have the ability to activate the inert gas by means of excitation, dissociation
and ionization and the species generated by this route, will easily go through other
reactions, producing new molecules. Hence the gas (CO2) does not need to be heated
completely. By this approach, endothermic reactions e.g. DRM or CO2 splitting can
take place at moderate reaction conditions with significant energy consumption. It is
not energy efficient to generate plasma by using electrical power [183].
Microwave plasmas (MW), gliding arc (GA), and dielectric barrier discharges
(DBD) are the most commonly used plasma for CO2 transformation. Microwave
plasma was reported to be most efficient but under peculiar conditions, which would
be appealing for commercialization, produces a drastic drop in efficiency[183, 184].
The energy efficiency of GA plasma is also comparatively high at atmospheric
pressure. Its energy efficiency was 60% for conversion of up to 16% (DRM) and 43%
for conversion of 18% (CO2 splitting). DBD plasma is not much energy efficient (2-
10%), but as indicated for different other applications, energy efficiency can be
improved using a packing of dielectric in the reactor. Furthermore, its design is very
simple and works at atmospheric pressure, which is highly useful for upscaling. It has
already been shown for the generation of ozone on large-scale and therefore, it has
huge potential for commercialization.
Plasma is also considered the fourth state of matter because matter converts
into solid, liquid, gas and finally plasma and Irving Langmuir first introduced this
term in 1928. Plasma is an ionized state of matter having at least one unbound
electron, making ions with positive charges. Practically, the degree of ionization of
plasma varies from 100% (completely ionized) to 10-6 (partially ionized). The
presence of excited species in plasma leads to light emission. Plasmas can be found in
many applications such as microelectronics, material science, environmental and
medical. They are also used as sources of light due to their ability of emitting light.
Other than natural plasma, it can be classified into fusion plasmas (high temperature)
which are actually completely ionized and gas dischargers (weakly ionized). Another
classification can be made on the basis of thermal equilibrium. The plasma
temperature can be determined by energies of several species present in plasma and
their different kinds of motion (rotational, translational and vibrational) [51].

4.4.1 Different plasma technologies for methane – carbon dioxide reforming


Plasma utilized for methane – carbon dioxide reforming can be categorized
into thermodynamically equilibrium (thermal plasma) and thermodynamically non-
equilibrium plasmas (cold plasma). In cold plasma, the electrons as compared to
other species possess more kinetic energies and plasma average temperature used to
be very close to room temperature. The temperature of electrons and other species is
approximately the same (having a range of 1000 K). Since electron density is higher
in thermal plasma, because of the large number of collisions between other species
and electrons, the energy of electron obtained from the electric field consumes in the
heating of other species to reach thermodynamic equilibrium in both electrons and
other species. Thermal plasma has thermo-chemical and electron-induced chemical
reactions.

4.4.2 Reforming of methane-carbon dioxide by cold plasma


Cold plasmas are generally heterogeneous inside the discharge space resulting
in small reaction region which restricts treatment and conversion capacity. The
catalytic process is introduced in the reforming of methane-carbon dioxide by cold
plasma in order to increase the rate of conversion. In cold plasmas, a catalyst can be
introduced either by placing it (catalyst) in the discharge space or at the end of
discharge space [185]. Different types of plasmas have been used to reform methane-
carbon dioxide and their characteristics have been included below. At the end of the
section, Table 7 is added which presents a comparison in terms of conversion rate and
products obtained between types of plasma.
4.4.2.1 Microwave discharge
The reactor of microwave plasms incorporates waveguides, microwave
generator and resonance cavity placed into a quartz tube whereby chemical reactions
are carried out to produce plasma. This plasma generally operates at a wide range of
pressure and frequencies (GHz). MW discharge exhibits excellent discharge
uniformity and has greater discharger space as compared to other cold plasmas at
atmospheric pressure. The temperature of particle and electron is approximately 2000
K and 0.4 – 0.6 eV, respectively in MW discharge [186]. The reforming of methane-
carbon dioxide through MW discharge yields acetylene, synthesis gas and ethylene.
The excessive amount of methane in feed leads to more hydrocarbons in the
conversion process. The reforming of methane-carbon dioxide through MW discharge.
The chemical reaction is induced by the high temperature of the gas and the high
energy of electrons. Despite all, the equipment used in MW discharge is bulky and
complex and it is problematic to acquire large equipment favorable for industrial
demands.

4.4.2.2 Corona discharge


It is heterogeneous discharge and produced at atmospheric pressure having a
high intensity of electric field for the decomposition or ionization of feed gas. Corona
discharge incorporates two (irregular) electrode; high curvature and low curvature
(shown in Fig.15). This discharge is formed as the electric break down occurs in the
vicinity of high -curvature electrode when the applied voltage surpasses the given
limit. The temperature of gas and electron is approximately 400 K and 3.5 – 5 eV,
respectively in corona discharge [187]. The high density of electrons (1015 – 1019 m-
3) mostly surrounds the high-curvature electrode. In corona discharge, reaction
volume is restricted by localized breakdown which makes this discharge difficult to
obtain a high treatment capacity in the reforming of methane-carbon dioxide by this
discharge.
Line
electrode
Power
supply Tube
electrode

Discharge
region

Fig.15. Schematic diagram of the corona discharge reactor [185].


4.4.2.3 Gliding arc discharge
There are several diverging electrodes made up of metals in the reactor and arc
discharge is produced when high voltage is exerted on the electrode (Fig.16). Lesueur
et al. patented the principle of this discharge in 1988 and it was further improved by
Czernichowski et al. [188]. This discharge is eminent by arc flame and primarily
produces on the nearest gap, that arc is further extended and its root slides along the
surface of the electrode by drawing of the rapid flow of employed gas until quenches
plasmas downstream. Based upon the flow rate of gas and discharge power, the GAD
can be classified into cold or thermal plasma [189].

4.4.2.4 Dielectric barrier discharge (DBD)


It was first discussed by Siemens and is also known as silent discharge [190,
191]. It is mainly used to produce ozone and typically incorporates planar electrodes
which are positioned parallelly having a gap of few millimeters and one of these
electrodes has a dielectric cover which prevents the generation of arc and spark
(Fig.17).
Mixing
chamber

Electric heat
Pure methane
tube
cylinder

Control B
Pure carbon panel
dioxide cylinder Electric
thermocouple

Nozzle

Voltage regulator

Teflon plate

AC power
supply Transformer
Quartz
encloser
Electric
Electrode thermocouple

Mass flow
meter

FT-IR

Hydrogen sensor

Outlet

Fig.16. Setup of gliding arc discharge [185].


High voltage
electrode

Product

Power
supply Grounded
electrode

Quartz tube

Gas
Feed
distributor

Fig.17. Schematic diagram of the reactor of BDB [185].


The material of dielectric could be ceramics, glass, polymer, quartz or any
material of small dielectric loss and high strength of breakdown [187]. At atmospheric
pressure, DBD is also a heterogeneous discharge and needs alternating voltages for
operation. The temperature of gas and electron is approximately in hundreds K and 1
– 10 eV, respectively in DBD [192]. This discharge was utilized for the reforming of
methane-carbon dioxide by many researchers (Table 6). DBD shows lower
conversion rates in most situations and syngas selectivity are low and products are
normally complex. The reaction performances are not impressive even if the
combination of plasma and catalyst is used. The intrinsic features of DBD cause more
production of hydrocarbons in the reforming of methane-carbon dioxide. In the filar
channel, the CH4 molecules are easily decomposed into radicals of CHx with high
density and energy of the electron. These radicals quickly leave the channel and move
into other region (free-discharge zone) where they interact to produce hydrocarbons
[185].

4.4.2.5 Atmospheric pressure glow discharge (APGD)


This discharge is like an arc and produced by employing some hundred volts
between electrodes [193] using a ballast circuit to prevent the conversion of glow into
the arc (Fig.18).
Mixer

Cold plasma
jet reactor
Power
supply

CH4 CO2 N2
Catalyst bed

Cool trap

Fig.18. APGD schematic diagram of the experimental setup [185].


The ‘glow’ demonstrates that the discharge plasma is luminous contrary to the
dark discharge [181]. Glow discharge can work over a long range of pressure. The
electron density is 1018–1019 m-3 and the temperature of gas and electron is
approximately in 2000 K and 1 – 2 eV, respectively in APGD [194, 195]. The
efficiency of energy conversion through the jet of cold plasma is increased
considerably as compared to other types of cold plasmas. This is obliged to the mode
of plasma utilized and design configurations of reactors. An AC jet of cold plasma
with a frequency of 20 kHz has high plasma temperature and density of electron as
compared to DBD and corona discharge. Such high values of plasma temperature
cause catalyst heat up and consequently increase the synergistic effect of catalyst and
plasma, similarly, high density of electron directly enhances the reforming reaction.
APGD is a promising form of plasma for methane-carbon dioxide reforming
attributable to high density and energy of electron plus appropriate plasma
temperature. However, the enlargement of this process on a commercial scale is one
of the notable challenges.

4.4.3 Thermal plasma


It is a uniform and continuous plasma produced in the pathway of an electric arc. It is
characterized by a high density of electron (1019 – 1020 m-3), a high temperature of
electron and has distinct chemical and thermal effects. It has numerous applications
and is widely adopted on the industrial scale [63]. It has several types, shown below
(Fig. 19). Thermal plasma has a high potential for reforming of methane-carbon
dioxide due to chemically reactive elements and high temperature. In comparison with
the aforementioned cold plasmas, reforming of methane-carbon dioxide using single-
anode TM shows many advantages in terms of conversion efficiency, selectivity, by-
products, and treatment capacity.

Thermal plasma

Alternating Radiofrequency
Direct current High frequency
current arc inductively
arc torch capactive torch
torch coupled torch
Fig.19. Types of thermal plasmas.
Moreover, in thermal plasma input gas is directly injected into the plasma jet
instead of the discharge region between the two electrodes. The conversion efficiency
and treatment capacity can be increased if feed gases are directly added in the
discharge region as plasma producing gas (Fig.20).

1.0
1.0
CO2
0.9
0.9 CH4
0.8
Conversion

Selectivity

0.8 CO
0.7 H2
0.7
0.6
0.6
0.5
0.5
0.4
0.15 0.20 0.25 0.30 0.35 0.15 0.20 0.25 0.30 0.35
Փ m3/h/kW) Փ m3/h/kW)
1.00
1.00
0.95
0.95
0.90
Conversion

Selectivity
CO2
CH4 H2
0.90 0.85 CO

0.80
0.85
0.75
0.80 0.70
0.12 0.13 0.14 0.15 0.16 0.12 0.13 0.14 0.15 0.16
Փ m3/h/kW) Փ m3/h/kW)
Fig.20. Experiments performed by ref. [185] to compare the conversion and selectivity for
thermal plasma.
The literature cited in this review (Table 6) shows that thermal plasmas exhibit
higher selectivity, conversion, treatment capacity, and specific energy. It is quite
appropriate to develop this process for industrial applications.

4.4.4 Hydrogenation of CO2


Owing to the expensive hydrogen, there is limited research on hydrogenation
of CO2 through plasma. Economically viable and sustainable production of H2 is
essential for approaches depending on the utilization of H2 as co-reactant. However,
CO2 hydrogenation through plasma has recently attained a lot of attention.
Table 6. Different types of plasma along with their features.
Types of Catalyst H2/CO2 Conversion Major products Reference
plasma of CO2 (%)
DBD CuO/ZnO/Al2O3 3:1 13.3 methane, methanol [196]
DBD Ni/zeolite 4:1 96 CO, methane [197]
packed-bed Ni-CexZr1-xO2 4:1 78 methane [198]
DBD
DBD Cu/g-Al2O3, - 7.5 CO, H2O, and CH4 [199]
Mn/g-Al2O3 and Cu-
Mn/g-Al2O3
DBD - 1:9– 9:1 2–7 CO, H2O and CH4 [200]
MW - 1:1 - CO, acetylene, [201]
discharge methanol, ethylene,
RF - 4:1 26 CO and H2O, as [202]
discharge well as
CH4
MW - 3:1 82 CO, H2O, C2H4 [203]
MW Ni/TiO2 1:9 23 CO [204]
Surface - 1:1 15 CO [205]
discharge

It can be concluded that CO2 hydrogenation through plasma exhibits poor


performance, even if we compare it with CO2 splitting and DRM. To date, this
approach is not much successful because the combination of hydrogen and carbon
dioxide is not favorable for the production of methanol in the single-step process.
4.4.5 Future perspectives
Different plasma approaches for CO2 transformation have been discussed in
the above sections. Plasma technology has a high potential for CO2 processing due to
high temperature, enthalpy content, electron density, and treatment capacity. However,
there are several challenges that restrict its wide-scale applications.
A standardized framework is highly needed in data reporting to enable data
comparison inside and outside the plasma technology. The important standard here is
to compare the energy efficiency and conversions. As far as conversions are
concerned, it needs clear data for diluting agents when utilized, because these agents
significantly affect the plasma process. Moreover, the impact of gas expansion in
plasma operation, which is neglected, but can tremendously affect data accuracy e.g.,
the conversions acquired via GC measurements. Regarding energy efficiency, at first,
there should be no interpretation space about the reported power (whether it is
measured or applied power) and through which approach this power is acquired.
Secondly, there should be an aim to ascertain the important products along with their
selectivities, to measure the energy efficiency depending upon the reaction enthalpy
including fuel generating efficiencies. Thirdly, the literature on power supplies and
corresponding efficiencies and optimization is urgently needed to compare conversion
efficiencies (solar to fuel). Generally, for materials testing for plasma catalysis,
reporting data as well as a framework for experimentation are quite essential for the
development of plasma technology which will highly assist the plasma community
and CO2 transformation community.
Energy efficiency in terms of SEI, it is obvious that mostly employed SEIs are
extremely high, especially when used with catalysts. To achieve highly efficient
energy conversions, low SEIs (0.1–5 eV) should be targeted. Additionally, because
plasma technology is more susceptible to feed impurities, more investigations and
researches are required towards the impact of gas compositions on the chemical
(noxious by-products, distribution of products) and physical side.
Table 7. Different plasmas with percentage of obtained products at different power.

Type of plasma Feed flux CH4 Power Conversion Selectivity (%) H2/CO SE References
(mL/mi) /CO2 (W) (%) (kJ/mol)
CH4 CO2 CO H2 Acetylene Ethylene (C2H4)
(C2H2)
Cold plasmas
MW discharge 200 3/2 30 70.8 68.8 75 - 17.8 4.1 1.5 307 [206]
DC Corona 60 1/2 63 94.1 77.9 97.1 69.4 - - 0.6 1134 [207]
Discharge
Corona discharge 43 1/1 46.3 62.4 47.8 66.8 70 15.8 1.5 1.2 1798 [208]
Gliding arc discharge 1000 1/1 190 40 31 62 50 12 - 0.9 608 [209]
DBD and Ni/Al2O3 30 1/1 130 55.71 33.48 60.9 51.92 10.12 - 1.0 10385 [210]
DBD 20 1/1 107.4 72.8 44.4 82 70 - - 1.0 7289 [211]
Cold plasma jet and 8300 4/6 770 60.06 40.35 96.79 96.87 - - 1.0 134 [212]
Ni/Al2O3
Thermal plasmas
Bioanode thermal 73300 4/6 18000 78.71 64.80 96.79 28.85 / / 0.8 274 [213]
plasma
Single anode thermal 21700 4/6 8500 87.98 84.34 82.27 43.48 / / 0.7 520 [214]
plasma (H2)
Single anode thermal 33300 4/6 9600 92.32 82.19 90.15 75.43 / / 0.7 290 [215]
plasma (N2 + Al2O3)
4.5 Solar thermochemical conversion
Traditional industrial processes rely on hydrocarbons for producing fuels and
metals and they are highly carbon and energy-intensive. The depletion of fossil fuels
and environmental changes has led to the development of solar industrial processes.
The concentrated solar approach provides the opportunity of transforming solar
radiation into electrical, chemical and thermal forms. On the other hand, the
conventionally concentrated approach incorporates the utilization of solar energy to
drive many chemical reactions [216] in several applications. Solar thermochemical
operations utilize solar energy to drive highly endothermic reactions.
Attempts were directed to switch dependency from fossil fuels to hydrogen
economy after the fuel crisis of 1973. Consequently, studies on thermochemical
decomposition of water to produce hydrogen were initiated. The oil embargo and high
prices of oil ensured the momentum of investigations on thermochemical processes
for the next decade. However, the number of researches declined up to the year 2000
due to the stable oil market. With an emphasis on environmental changes, solar
thermochemical processes again received attention in the mid of the year 2000 and the
investigations have been increasing.
Concentrated
solar energy

1st Step: Solar reduction O2

MOox  MOred + O2
MOox MOred
2nd step: Oxidation

H2O MOred + H2O  MOox + H2 H2


CO2 MOred + CO2  MOox + CO CO

Fig. 21. Multi-step CO2 dissociation [51].


4.5.1 Working principle
With the help of solar energy, CO2 can be reduced in many ways. The most
efficient method is the direct uses of sunlight because it does not require any
additional energy and has no negative impact on the atmosphere [57]. The further two
types of this method are thermal conversion and quantum conversion. In the former
type, firstly solar radiation is absorbed and then work is extracted. The latter type
includes extraction of work directly from the semiconductor or any other compound
which are used as light absorber [58]. As mentioned before, concentrated solar energy
in the solar thermochemical process is utilized as high-temperature heat to drive high
endothermic reactions. Thermal dissociation of CO2 which involves only a single step
requires high temperatures and separation techniques to prevent the formation of an
explosive mixture (CO and O2) [58,59]. In the multi-step thermochemical conversion
of CO2, metal oxides are used.
Multi-step CO2 dissociation employing metal oxides in redox reactions detour
the issue of separation and additionally, they permit operation even at moderate
temperatures. In the first step (endothermic), solar thermal reduction of MOox to M or
MOred is carried out and the second step (exothermic) involves oxidation of MOred
with CO2 to produce CO (Fig. 21), enabling the MOox to be used again in the first step
[217]. M denotes metal e.g., Zn, Fe or Ce.
The direct utilization of sunlight in solar thermochemical technology is the
main advantage. The concentrating solar technique is currently used for electricity
generation and can be combined with high-temperature reactors to generate solar fuels
on a wide-scale at a cheap cost and to attain high solar-to-fuel efficiencies [58].
Currently, solar flux ratios (> 2 MW m-2) can be attained using a solar dish and tower
systems. A comprehensive review related to solar concentrating approaches can be
found in [218]. The solar thermochemical cycles have the ability to achieve a greater
efficiency compared to those methods which utilize energy vector or a small portion
of the solar-spectrum and are simpler in design [62,64]. This ability is due to the
phenomenon that solar-based thermal processes work at high temperatures and use the
complete solar spectrum and offer a thermodynamically suitable way to produce solar
fuels [51].

4.5.2 Categorization of solar thermochemical processes


In Fig. 22, different types of solar thermochemical processes are given.
Broadly, this process has two types: the production of industrial products and fuel
generation. Many industrial products (aluminum, zinc, lime, carbon nanotubes) can be
produced by solar thermochemical systems and it has a high potential.
Solar
thermochemical
process
Syngas/
Industrial
Hydrogen
applications
production

Recycling of
waste and
utilization
Solar Hydrocarbon Thermochemical
thermolysis feed cycles
Carbide,
Nitrides

Metal oxide Multistep cycles


Ammonia
Gaseous feed Solid feed
production

Carbon Sulphur Iodine


Methane Hybrid cycles Pure thermal
nanotubes

Solar processes Thermal Calcium


Reforming
materials decomposition Hybrid sulphur bromine cycle
Copper chloride
cycles

Lime Steam
production reforming

Dry reforming

Fig.22. Different types of solar-based thermochemical processes.


The up-gradation of carbonaceous feed yields synthesis gas and solar fuels
such as hydrogen can be generated by solar thermolysis incorporated in a
thermochemical cycle. Synthesis gas can also be produced by the combined splitting
of water and carbon dioxide in the thermochemical cycle. The water, accompanied by
process heat, is directly dissociated at high temperatures (1900–4000 K) in solar
thermolysis. Solar energy can be utilized for gasification and reforming of methane,
biomass, and coal that incarnate solar energy in the produced gases. Thermochemical
cycles are actually a chain of chemical reactions that decompose carbon dioxide and
water to generate carbon monoxide or hydrogen respectively. Contrary to solar
thermolysis, the chemical reactions in the thermochemical cycle are carried out at low
temperatures (800–2100 K). Thermochemical cycles are categorized as metal oxide
reduction-oxidation based cycles and multi-step cycles. In metal oxide-based cycles,
hydrogen and oxygen are produced by periodic oxidation and reduction of redox
materials. Depending upon the material, this type of thermochemical cycle can be
further categorized into volatile and non-volatile cycles. While multi-step cycles
involve compounds of chlorine, bromine, and sulphur, etc., and these elements
undergo a chain of reactions to decompose water for oxygen and hydrogen production
in different steps of the reaction. Based on the electrolysis, further classification of
multi-step cycles is hybrid or pure thermal cycles.

4.5.3 Reactors
Reactors in solar thermochemical technology are of extreme importance and
have attained a lot of research interest [219,220]. There are two general types of
reactors utilized for experimentation of solar thermochemical systems: directly or
indirectly heated. The reactants are heated by the direct solar radiation in the former
type, while the non-transparent surface is used to receive the solar energy and then
this heat is transferred to the reactants in the latter type. Depending upon the working
principle, these reactors have further types and few reactors have been tested in pilot-
scale plants and laboratories. Further classifications of directly heated reactors are
volumetric reactors and in this type of reactors, catalytic absorber directly absorbs the
solar energy and further transport it to feed gas. Depending upon the absorber,
honeycomb, structured and foam are other types of volumetric reactors.
Comparatively, foam-based reactors have a much better mechanism of heat transfer
which avoids flow instabilities during working at high temperatures and these reactors
exhibit more porosity that permits solar energy to enter deeper in the material [221].
The indirectly heated reactors incorporate molten salts, sodium, solid particles
or air as heat transfer medium. For methane reforming, solid particles were used at
NREL. The reactor used in this process works at 1870–2120 K and generates some
carbon. In the Asterix project [222,223], the air was used as the heat transfer medium,
this mechanism allows reforming design independent of solar considerations and the
influence of any variable such as clouds can be reduced by using thermal storage.
MoSTAR for methane reforming has been developed in Japan and primarily used to
obviate the cloud effects [224]. After successfully employing at laboratory scale, they
are extending it to use on five kilowatts dish concentrator. In another project, molten
salts were used to perform the solar reforming of methane in CoMETHy [225].
On the other hand, there is a main challenge of coking in other reactors.
Carbon particles increase radiative heat transfer in the aerosol reactor. Another
advantage is working at a high temperature which means higher conversion rates that
directs to more products and also avoids catalyst material [226]. The attack of CO2 on
reactor walls can be avoided by maintaining the ratio of methane to carbon dioxide
(>1). Tubular reformers with the different number of coils have been used for
methane reforming in Australia. The prime focus of these reformers was to conduct
methane reforming at relatively low temperatures (820–970 K). The process at low
temperature reduces heat deficiencies from the reactor by using the maximum amount
of solar energy. Table 8 presents the reactor performance and test conditions for
different experimental setups.

Table 8. Different types of reactors with their test conditions and obtained conversions.
Type of reactor T (K) P Qsolar Reforming Conversion Efficiency Reference
(bar) (kW) element (%) (%)
volumetric 970- 3.5 300 CO2 80 - [227,228]
1130
volumetric 1173 15 400 steam 80 80 [229–231]
Tubular reformer 1120 20 25, 500, Steam/CO2 - - [232,233]
1200
volumetric 1473 1 150 CO2 70 85 [234,235]
volumetric 1473 4-9 40 CO2 97 79 [236,237]
volumetric 1033 8.5 220 steam - - [238–240]
Tubular receiver 1273 7.8 170 steam 93 - [222,223]
Sodium reflex 958- 2-6 1.5-7.8 CO2 50-70 - [241,242]
1098
Aerosol 2073 - 10 CO2 70 - [226]
Tubular absorber 1193 1 5 CO2 90 26 [224]

4.5.4 Metal oxide redox pairs


There are many renowned redox cycles which can dissociate H2O and CO2
and have theoretical efficiencies more than 40%. Generally, volatile and non-volatile
are two-cycle classes considered for metal oxides. Doping techniques are frequently
used in both types of non-volatile cycles to improve their physical, thermodynamic
and kinetic characteristics. The commonly employed oxide pairs with reaction
schemes are summarized in Table 9.
Table 9. Commonly used redox pairs [243].
Type Cycle Symbol Reaction scheme
Volatile Tin Oxide SnO2 SnO 2 (s)  SnO(g)
Zinc ZnO
Oxide ZnO(s)  ZnO(g)
Non-volatile Ferrite MxFe3-xO4 M x Fe3 x O 4  xMO   3  x  FeO
(stoichiometric) Iron oxide Fe3O4
Hercynite Fe3O4+3Al2O3 Fe3O 4  FeO
Fe3O 4 +3Al2 O3  3FeAl2 O 4
Non-volatile Perovskite ABO3 ABO3  ABO3
(stoichiometric) Ceria CeO2
Doped MxCe1-xO2 CeO2  CeO2
Ceria
M x Ce1 x O2  M x Ce1 x O2

Types of two step


cycles
Volatile cycles Non-volatile cycles
 Consist of metal oxides that  Utilize metal oxides which
undergo gas–solid phase remain in the solid state during
transitions reduction.
 Have a greater oxygen exchange  Poor oxygen exchange capability
capability than non-volatile  Thermodynamically unfavorable
reactions reduction
 Thermodynamically more  No issue of products
favorable reduction due to the recombination
increased entropy
 Products must be quenched
rapidly to avoid recombination
Non-stoichiometric cycles Stoichiometric cycle
remain crystallographically stable generally form solid solutions upon
while the lattice accommodates reduction. They have a greater
changes in anion or cation vacancies oxygen storage capacity than non-
concentrations stoichiometric reactions

Fig.23. Classification of two-step cycles.


4.5.4.1 Perovskites
Perovskites are represented by the high yield of oxygen at relatively high
temperatures and they also incorporate energy-demanding reduction-oxidation. Their
redox properties are determined by the cation which is present at the M site. It is a
category of nonstoichiometric oxides and denoted by ABO3-δ. These oxides are
commonly unexplored for the metal reduction in thermochemical cycles. A and B are
cation sites on which dopants can be replaced and hence, material configurations are
significantly larger in perovskites compared to ceria. A thermodynamic study
involving the extraction of entropies and enthalpies and data evaluation of oxygen
non-stoichiometry has demonstrated that LSM perovskites denoted as La1-xSrxMnO3-δ
can enhance the capacity of oxygen exchange than pure ceria [244]. As shown by
experiments, though reduction extents are significantly greater, oxidation is
thermodynamically less feasible leading to incomplete oxidation. Nonetheless, overall
CO production from the decomposition of CO2 remains substantially higher compared
to ceria. McDaniel et al. explored a more favorable and promising type of perovskites.
They doped Sr2+ and LaAlO3-δ with Mn on A and B sites, respectively [245]. They
conducted oxidization at 1273 K, reduction at 1253 K and produced six and nine
times more CO and H2, respectively than ceria (Fig. 23). La1-xSrxFeO3-δ for looping
applications has been studied by [246]. They produced syngas by using methane in
the reduction process which lowered the temperatures (1000–1273 K). Large amounts
of hydrogen were produced when a mixture of La0.3Sr0.7FeO3- δ and 5% NiO were
used. To date, Lanthanum manganites – which uses Mn4+/Mn3+ ions for redox
reactions – is the family of perovskites which are extensively considered and studied
[244, 247-250]. These materials of perovskite exhibit low oxidation performance and
need large quantities of oxidizing gas that significantly increases heat losses [251,
252]. Few perovskites material possesses high molar mass as compared to ceria and
heat capacity [244, 253]. For a specific temperature difference in two steps, the high
heat capacity of these perovskites results in a high loss in terms of sensible heat.
These studies demonstrate that, based on different doping schemes, many perovskites
can be developed which can efficiently increase the yield of desired products at low
temperatures. The development of such materials which can be transformed below
1273 K can be a turning point for the production of solar fuels [254].

1.4 1.2
(a) (b)
1.2 1.0
Production of CO
Porduction of H2

1.0
CeO2 0.8 CeO2
0.8 SLMA3
0.6 SLMA3
SLMA2
0.6 SLMA1
SLMA2
0.4 SLMA3
0.4
0.2 0.2
0.0 0.0
0 200 400 600 800 0 200 400 600 800
Time (s) Time (s)

Fig.24. Rates of H2 (a) and CO (b) production in terms of time for different perovskites.
SLMA1: Sr0.6La0.4Mn0.6Al0.4O3-δ; SLMA2: Sr0.4La0.6Mn0.6Al0.4O3-δ; SLMA3: Sr0.4La0.6Mn0.4Al0.6O3-δ
[245].
4.5.4.2 Ceria
Flamant and Abanades conducted an experiment in which CeO2 was reduced to Ce2O3
at 2273 K. They observed extensive sublimation (higher than 50%) [255]. Based on
their experiments, Haile and Chueh developed a cycle to reduce CeO2 into CeO2- δ
which operates at low temperatures for stoichiometric reduction [256]. Chueh et al.
tested and established a solar reactor based on ceria-materials and successfully
demonstrated the cycle viability under real solar conditions [257]. Panlener et al.
investigated defect chemistry of ceria and oxygen non-stoichiometry at high
temperatures. They concluded that iron oxide-based cycles have a high capability of
oxygen exchange compared to ceria and sintering is not as such intricate due to a high
melting point [258, 259]. Hence, it is not essential to dissolve or support them in more
strong structures. The decomposition of water and carbon dioxide was experimentally
demonstrated in [260], producing syngas by changing the molar ratio of water and
carbon dioxide in the reactor. In pure ceria, ionic and electronic conductivities directly
affect diffusion rates [261].

Heat transfer limits the reaction extents in reduction more than surface kinetics
or oxygen diffusion. CO2 oxidation rates depend on surface area and are highly rapid
than Fe-based materials of smaller lengths [260,263]. Haile and Chueh studied the
cyclical stability and found that grain growth and sintering decrease the yield
significantly in early (100) cycles, however, the production rates of hydrogen and
stability of hydrogen/oxygen was remarkable for the next 400 cycles [264]. Although
ceria has an attraction in terms of stability and kinetics, its storage capacity is low
than other materials. In Fig. 24(a), production of oxygen due to CeO2 reduction with
respect to Poxygen and temperature are demonstrated. The reduction is generally 1/8 of
Co-Fe at T = 1473 K, x = 0.15, Poxygen = 10-10 bar is shown in Fig. 24(b). Efficient heat
restoration is necessary for this cycle to achieve economic feasibility and solar-to-fuel
efficiency [265].

4.5.4.3 Ferrites
Replacement of transitional metal in the lattice of magnetite can enhance the
oxygen storage-capacity than pure magnetite. An increase of metal concentration in
the lattice can increase the reduction extent at given conditions (Fig. 25). Ferrites have
been studied thermodynamically and tested experimentally using metal cations
CoFe2O4 [266], ZnFe2O4 [267], MnFe2O4 [268], NiFe2O4 [269], Mn0.5Zn0.5Fe2O4 [270]
and Ni0.5Mn0.5Fe2O4 [271] etc. on A site [272]. The testing of ferrite-powders has
determined the impact of different parameters on the production of hydrogen and
oxygen and the temperature values suitable for cyclic operations [273]. The major
drawback is a high reduction temperature (1500–1700 K) which results in oxide
sintering. Much attention has been devoted to addressing this issue by utilizing redox
reagent at elevated temperatures. Ferrites are relatively more stabilized when
dissolved in chemically inert materials to avoid melting and sintering. Kodama et al.
conducted experiments incorporating Co and Ni ferrite with ZrO2 or YSZ [22,48].
The production of hydrogen was high than Fe-YSZ process, as demonstrated by
Allendorf et al. [274].
Relating to the composition of materials, it can be said that ferrite containing
Mn and Zn demonstrates problems related to phase stability and volatilization,
respectively [275]. The robust ferrites are CoFe2O4 and NiFe2O4 which operate
efficiently and reliably and at real-world conditions. Concerning to reaction
mechanism, data obtained from mass-spectrometers and thermogravimetric analyzers
have been utilized by several research groups for the establishment of kinetic models
and their subsequent extraction [276-278]. Interestingly, the latest studies have only
converged commonly practiced patterns. Thermal decomposition of carbon dioxide
and water using ferrites is fast compared to their typical reduction and hence
consumes less time.

4.5.5 Future perspectives


Solar thermochemistry has evolved as a feasible route that uses concentrated
solar energy and is currently being used on a commercial scale for power generation.
Even though significant progress has been achieved in the solar thermochemical
transformation of CO2 utilizing metal oxides, a lack of basic research to investigate
the performance of metal oxides subjected to high temperatures in redox cycles has
impeded the development of new materials. Basic problems coupled with surface
chemistry, production methods of material, oxygen transport, the impact of
thermochemical cycling on constituents and structural changes still require efficient
solutions. Moreover, thermal losses in the reactor largely limit the cycling rates and
efficiency which results in the poor transfer of radiative and conductive heat across
the structures of metal oxides.
It is thermodynamically favorable to convert CO2 through redox reactions but
the conclusive factor determining the commercial feasibility is solar-to-fuel efficiency
and currently, efficiencies more than 10% still need experimentation with scalable and
robust solar reactors. It is essential to discover new materials that have large
capabilities of oxygen exchange at average temperatures and their execution in solar
reactors. Furthermore, material stability and fast chemical kinetics over hundreds of
cycles should be demonstrated for every material considered. This is an important
factor to get market acceptance because materials should stay active for hundreds of
cycles to hamper high costs associating with repeated replacements. Commercial-
scale usage of this technology is again determined by exploring earth-abundant
materials that can work at lower temperatures to attain high system efficiency.

5 Summary
It is evident from the above discussion that, irrespective of maturity,
commercialization and other detrimental aspects, all the current approaches have
particular advantages. Table 10 has been included to demonstrate the comparison
between all these technologies.
Table 10. Comparison of different technologies used for CO2 conversion.
Use of earth

Conversion
Renewable

Investment
and yields

Operating
Technology

Turnkey
Process
energy
metals

flexibility
cost

cost
Products

Traditional Yes - No High Yes Yes High Low


Photochemical Yes Direct Yes Low Yes Low Low Low
Biochemical No Direct No Medium Yes H/L High Low
Electrochemical Yes Indirect No High Yes Low Low Medium
Plasma chemical No Indirect Yes High Yes High Low High
Solar Yes Direct Yes High No High Low High
Thermochemical
*The reference [3, 51, 111, 112, 127, 181, 183, 243, 273] has been considered for
establishing this table.

The rare utilization of earth-abundant metals is one of the distinct


disadvantages of almost all the current adopted technologies, in terms of cost and
sustainability. A great challenge for all the aforementioned approaches to become
more efficient is to use easily accessible inexpensive materials. The discussion about
dependence on renewable energy (directly or indirectly) is subtle. The direct
utilization of renewable energy gives an edge to photochemical, biochemical and solar
thermochemical over other technologies that use indirect renewable energy. There is
no extra step of energy conversion in photochemical, biochemical and solar
thermochemical approaches and hence energy losses are prevented. In plasma and
electrochemical conversions, the efficiency of renewable generation of electricity
determines the overall energy efficiency of these approaches. However, it also shows
that these approaches (plasma and electrochemical) can depend on other sources of
renewable energy such as wave, hydro, tidal and wind power, enhancing their
employability, because they can be operated anywhere without the availability of
solar energy. The combined use of photo and biochemical technologies with artificial
lighting is also an indirect source of renewable energy. The wide-scale employability
of these sources is a challenge for easy transport and efficient storage of the electricity
generated. Thus, a stabilized grid is required which needs those technologies which
flexibly follow the periodic and sometimes inconsistent supply of electricity.
Approaches that can harness this type of energy and transform it into carbon-free fuels
could be crucial for the future infrastructure of energy. Most technologies (except
biochemical and photochemical) exhibit high conversion rates and productions and
solar to fuel efficiencies of these approaches are demonstrated in Table 10. Another
aspect is the separation of products when all the feed is transformed. Based on the
process, electrochemical and solar thermochemical CO2 transformations are the only
two approaches capable of producing separated products. For instance, pure splitting
of CO2 yields CO and oxygen streams. It is a notable advantage because their
separation (carbon dioxide and oxygen) is highly energy-intensive. Although, it is
expected that the use of membranes will provide a convenient and energy-efficient
way of separation. The products in plasma technology are in one feed and their
separation is possibly the biggest drawback of plasma technology. The research on
this topic needs more attention.
The direct oxidative route to generate more valuable products is highly
favorable instead of current involving indirect method, which first generates carbon
monoxide and hydrogen and then these chemicals are further processed. All the
technologies - except solar thermochemical - are capable of producing many valuable
chemicals directly, though the amounts of chemicals are extremely low. It requires
insights and research to achieve more yields from these technologies.
Currently, the costs of investment and operation are low for most approaches.
Excepting, biochemical technology (where both investment and operating costs are
high depending upon the type of bioreactor) and solar thermochemical technology
which has a high cost of investment due to the mechanism of solar energy
concentration.

6 Conclusions
The conversion of CO2 into fuels and chemicals is an alternate approach to
overcome both fuel shortages and climate change. Several CO2 conversion
technologies: photochemical, biochemical, electrochemical, plasma-chemical and
solar thermochemical have been undertaken to evaluate their performances for CO2
conversion into value-added products. All these technologies have some challenges in
their practical application for CO2 capture and transformation. The materials and
procedures for the generation of industrial feedstock and solar fuels are disparate and
can’t easily interrelate. The photochemical transformation of CO2 seems to be an
emerging approach and though, it has a complex reduction mechanism, the objective
is to reduce the steps experienced during the reduction process. A few possible ways
to overcome this issue are the selection of proper semiconductors with a developed
hybrid structure, band edges and morphology designing. Electrochemical reduction of
CO2 is also a promising technology whereby the catalyst is viewed as an important
factor in regulating yield, reaction rates, and selectivity. Metal electrodes can be
scaled down quite easily with low Faradic efficiency, high overpotential, and low
current densities. The use of microbes in biological technology is continuously
expanding for CO2 conversion into value-added fuel and chemicals and several
microbial species have been found which efficiently contribute to the conversion
process. Different organisms can be genetically modified to increase the yield and
efficiency of the biochemical process. The productivity of biochemical processes can
be enhanced by optimal selection of microbes, physiological parameters, and structure
of bioreactor. Plasma-chemical is also extensively being used for the transformation
of CO2 through the cold plasma and thermal plasma. The latter type of plasma is
considered the best option for reforming of CH4-CO2 because of highly reactive
elements and temperature and has many advantages over cold plasma in terms of
selectivity, efficiency, treatment capacity, and by-products. The direct oxidative
generation of oxygenates through plasma-chemical conversion seems to be very
promising and need more investigations. Solar thermochemical technology has
undertaken many leaps forward and appeared as a feasible path to use concentrated
solar energy. The direct utilization of solar energy makes this technology highly
useful and advantageous over other conversion processes. Concentrated solar energy
is currently being employed commercially for power generation applications and
reduction of CO2 and H2O into CO and H2. Liquid hydrocarbons can be produced by
further processing the syngas. Though solar thermochemical splitting of carbon
dioxide and water through metal oxides are thermodynamically convenient, higher
ηsolar-to-fuel still requires experimentation with powerful solar reactors.
Acknowledgments
This work was supported by the National Natural Science Foundation of China
(51522601; 51950410590), China Postdoctoral Science Foundation Fund
(2019M651284), and Fundamental Research Funds for the Central Universities
(HIT.NSRIF.2020054).

Declaration of interests

The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.
References
[1] H. Balat, E. Kırtay, Int. J. Hydrogen Energy. 35 14) 2010) 7416-7426.
[2] S.C. Roy, O.K. Varghese, M. Paulose C.A. Grimes, ACS nano. 4 (3) (2010)
1259-1278.
[3] S. DasW.W. Daud, Rsc Advances. 4 (40) (2014) 20856-20893.
[4] L. Jaeglé, L. Steinberger, R.V. Martin K. Chance, Faraday Discuss. 130 (2005)
407-423.
[5] W.J. Ong, M.M. Gui, S.-P. Chai A.R. Mohamed, Rsc Advances. 3 (14) (2013)
4505-4509.
[6] J.L. Sawin, F. Sverrisson, K. Seyboth, R. Adib, H.E. Murdock, C. Lins, A.
Brown, S.E. Di Domenico, D. Kielmanowicz L.E. Williamson. (2016).
[7] L. Kong, Z. Jiang, T. Xiao, L. Lu, M.O. Jones P.P. Edwards, Chem. Commun.
47 (19) (2011) 5512-5514.
[8] M. Aresta, A. Dibenedetto A. Angelini, Chem. Rev. 114 (3) (2013) 1709-
1742.
[9] U. Bossel, Proceedings of the IEEE. 94 (10) (2006) 1826-1837.
[10] A. Goeppert, M. Czaun, J.-P. Jones, G.S. Prakash G.A. Olah, Chem. Soc. Rev.
43 (23) (2014) 7995-8048.
[11] G.A. Olah, A. Goeppert G.S. Prakash, J. of Org.c Chem. 74 (2) (2008) 487-
498.
[12] W. Wang, S. Wang, X. Ma J. Gong, Chem. Soc. Rev. 40 (7) (2011) 3703-
3727.
[13] R.K. Pachauri, M.R. Allen, V.R. Barros, J. Broome, W. Cramer, R. Christ, J.A.
Church, L. Clarke, Q. Dahe P. Dasgupta, Climate change 2014: synthesis
report. Contribution of Working Groups I, II and III to the fifth assessment
report of the Intergovernmental Panel on Climate Change, Ipcc, 2014.
[14] D. Töbelmann, T. Wendler, J. Clean. Prod. (2019) 118787.
[15] S. AdamsC. Nsiah, Sci. Total Environ. 693) (2019) 133288.
[16] S. Hernández, M.A. Farkhondehfal, F. Sastre, M. Makkee, G. Saracco N.
Russo, Green Chem. 19 (10) (2017) 2326-2346.
[17] J. Albo, M. Alvarez-Guerra, P. Castaño A. Irabien, Green Chem. 17 (4) (2015)
2304-2324.
[18] M.K. Lam, K.T. Lee, Int. J. of Greenh. Gas Con. 14) (2013) 169-176.
[19] Z.H. Lee, S. Sethupathi, K.T. Lee, S. Bhatia A.R. Mohamed, Renew. Sust.
Energ. Rev. 28 (2013) 71-81.
[20] Y. Wang, Q. Liu, J. Sun, J. Lei, Y. Ju H. Jin, Energy Convers. Manage. 133
(2017) 118-126.
[21] D.J. DarensbourgS.J. Wilson, Green Chem. 14 (10) (2012) 2665-2671.
[22] G. Fiorani, W. Guo A.W. Kleij, Green Chem. 17 (3) (2015) 1375-1389.
[23] M. North, R. Pasquale C. Young, Green Chem. 12 (9) (2010) 1514-1539.
[24] M. ArestaA. Dibenedetto, J. Brazi. Chem. Soc. 25 (12) (2014) 2215-2228.
[25] M.E. Boot-Handford, J.C. Abanades, E.J. Anthony, M.J. Blunt, S. Brandani, N.
Mac Dowell, J.R. Fernández, M.-C. Ferrari, R. Gross J.P. Hallett, Energy Env.
Sci. 7 (1) (2014) 130-189.
[26] A. Goeppert, M. Czaun, G.S. Prakash G.A. Olah, Energy Env. Sci. 5 (7)
(2012) 7833-7853.
[27] P. Markewitz, W. Kuckshinrichs, W. Leitner, J. Linssen, P. Zapp, R. Bongartz,
A. Schreiber T.E. Müller, Energy Env. Sci. 5 (6) (2012) 7281-7305.
[28] J.A. Martens, A. Bogaerts, N. De Kimpe, P.A. Jacobs, G.B. Marin, K. Rabaey,
M. Saeys S. Verhelst, ChemSusChem. 10 (6) (2017) 1039-1055.
[29] M. Mikkelsen, M. Jørgensen F.C. Krebs, Energy Env. Sci. 3 (1) (2010) 43-81.
[30] E.A. Quadrelli, G. Centi, J.L. Duplan S. Perathoner, Chem. Sus. Chem. 4 (9)
(2011) 1194-1215.
[31] M. Aresta, A. Dibenedetto A. Angelini, Philosophical Transactions of the
Royal Society A: Mathematical, Physical and Eng. Sci. 371 (1996) (2013)
20120111.
[32] B. Metz, O. Davidson, H. Coninck, M. Loos L. Meyer, Cambridge University
Press, 2005.
[33] F.A. Rahman, M.M.A. Aziz, R. Saidur, W.A.W.A. Bakar, M. Hainin, R.
Putrajaya N.A. Hassan, Renew. Sust. Energ. Rev. 71 (2017) 112-126.
[34] J. Mack, B. Endemann, Energy Policy. 38 (2) (2010) 735-743.
[35] S. Bouzalakos, M. Mercedes, Elsevier, 2010, pp. 1-24.
[36] J.H. Han, Y.C. Ahn, J.U. Lee I.B. Lee, Korean J. Chem. Eng. 29 (8) (2012)
975-984.
[37] G.A. Olah, A. Goeppert G.S.J.T.J.o.o.c. Prakash. 74 (2) (2008) 487-498.
[38] A.A. Olajire, Energy. 35 (6) (2010) 2610-2628.
[39] H. Yang, Z. Xu, M. Fan, R. Gupta, R.B. Slimane, A.E. Bland I. Wright, J. of
Environ. sci. 20 (1) (2008) 14-27.
[40] H. HerzogD. Golomb, Encyclopedia of energy. 1 (6562) (2004) 277-287.
[41] K.H. Kaggerud, O. Bolland T. Gundersen, App. Therm. Eng. 26 (13) (2006)
1345-1352.
[42] K. Damen, M. van Troost, A. Faaij W. Turkenburg, Prog. Energy Combust.
Sci. 32 (2) (2006) 215-246.
[43] D.Y. Leung, G. Caramanna M.M. Maroto-Valer, Renew. Sust. Energ. Rev. 39
(2014) 426-443.
[44] M. Kanniche, R. Gros-Bonnivard, P. Jaud, J. Valle-Marcos, J.-M. Amann C.
Bouallou, App. Therm. Eng. 30 (1) (2010) 53-62.
[45] A.S. Bhown, B.C. Freeman, Env. Sci. Tech. 45 (20) (2011) 8624-8632.
[46] H. ChalmersJ. Gibbins, Proceed. Instit. Mech. Eng., Part C: J. Mech. Eng. Sci.
224 (3) (2010) 505-518.
[47] M. Anwar, A. Fayyaz, N. Sohail, M. Khokhar, M. Baqar, W. Khan, K. Rasool,
M. Rehan A. Nizami, J. Env. Managem. 226) (2018) 131-144.
[48] W.M. Haynes, CRC handbook of chemistry and physics, CRC press, 2014.
[49] Z. Jiang, T. Xiao, V.á. Kuznetsov P.á. Edwards, Philosophical Transactions of
the Royal Society A: Math. Phy. Eng. Sci. 368 (1923) (2010) 3343-3364.
[50] H.J. Freund, M.W. Roberts, Surface Science Reports. 25 (8) (1996) 225-273.
[51] R. SnoeckxA. Bogaerts, Chem. Soc. Rev. 46 (19) (2017) 5805-5863.
[52] N. Itoh, M.A. Sanchez, W.C. Xu, K. Haraya M. Hongo, J. Membrane Sci. 77
(2-3) (1993) 245-253.
[53] Y. NigaraB. Cales, Bull. Chem. Soc. Jpn. 59 (6) (1986) 1997-2002.
[54] Y. Fan, J.Y. Ren, W. Onstot, J. Pasale, T.T. Tsotsis F.N. Egolfopoulos, Ind.
Eng. Chem. Res. 42 (12) (2003) 2618-2626.
[55] H. Jun, M. Careem A. Arof, Renew. Sust. Energy Rev. 22) (2013) 148-167.
[56] F. FisherH. Tropsch, Brennst.-Chem. 9) (1928).
[57] W. Lewis, E. Gilliland W.A. Reed, Ind. Eng. Chem. 41 (6) (1949) 1227-
1237.
[58] R. Reitmeier, K. Atwood, H. Bennett H. Baugh, Ind. Eng. Chem.. 40 (4)
(1948) 620-626.
[59] N.J. Rostrup, Google Patents, 1974.
[60] A.J. Brungs, A.P. York, J.B. Claridge, C. Márquez-Alvarez M.L. Green, Catal.
Lett. 70 (3-4) (2000) 117-122.
[61] S. Wang, G. Lu G.J. Millar, Energy Fuels. 10 (4) (1996) 896-904.
[62] M. Alvarez-Galvan, N. Mota, M. Ojeda, S. Rojas, R. Navarro J. Fierro, Catal.
Today. 171 (1) (2011) 15-23.
[63] M. Gharibi, F.T. Zangeneh, F. Yaripour S. Sahebdelfar, Applied Catalysis A:
General. 443) (2012) 8-26.
[64] V. Havran, M.P. Dudukovic C.S. Lo, Ind. Eng. Chem. Res. 50 (12) (2011)
7089-7100.
[65] J.H. Lunsford, Catal.Today. 63 (2-4) (2000) 165-174.
[66] J. Ross, A. Van Keulen, M. Hegarty K. Seshan, Catal. Today. 30 (1-3) (1996)
193-199.
[67] B. Vora, J.Q. Chen, A. Bozzano, B. Glover P. Barger, Catal. today. 141 (1-2)
(2009) 77-83.
[68] J. Zaman, Fuel Process.Tech. 58 (2-3) (1999) 61-81.
[69] S.G. Jadhav, P.D. Vaidya, B.M. Bhanage J.B. Joshi, Chem. Eng. Res. Des. 92
(11) (2014) 2557-2567.
[70] L. Torrente-Murciano, D. Mattia, M. Jones P. Plucinski, J. of CO2 Util. 6
(2014) 34-39.
[71] G.J.C. Prieto. Chem. Sus. Chem. 10 (6) (2017) 1056-1070.
[72] J.E.L. Russell R. Chianelli, G. Alexander Mills, Catal. Today. 22 (2) (1994)
361-396.
[73] A.N. Pour, J. Karimi, M. Housaindokht, M.J.R.K. Hashemian, Reac. Kinet.
Mech. Cat. 122 (1) (2017) 605-624.
[74] L.S.G. Yunnan, Z. Zhenqing, T. Hengcong, S. Zhenyu, J. Phy. Chem. 34 (8)
(2018) 858-872.
[75] S.N. Riduan, Y. Zhang, Dalton Trans. 39 (14) (2010) 3347-3357.
[76] J. Tollefson, Nat. 462 (7276) (2009) 966-967.
[77] J.N. ParkE.W. McFarland, J. Catal. 266 (1) (2009) 92-97.
[78] N. Shehzad, M. Tahir, K. Johari, T. Murugesan M. Hussain, J. CO2 Utilization.
26) (2018) 98-122.
[79] M. TahirN.S. Amin, Renew. Sust. Energy Rev. 25 (2013) 560-579.
[80] A. Pichon, Energy Mater. 2010.
[81] W. Tu, Y. Zhou Z. Zou, Adv. Mater. 26 (27) (2014) 4607-4626.
[82] M. Tahir, N.S. Amin, Energy Convers. Manage. 76) (2013) 194-214.
[83] H. Abdullah, M.M.R. Khan, H.R. Ong Z. Yaakob, J. CO2 Util. 22 (2017) 15-
32.
[84] Y. Zheng, L. Lin, B. Wang X. Wang, Angew. Chem. Int. 54 (44) (2015)
12868-12884.
[85] P. Niu, Y. Yang, C.Y. Jimmy, G. Liu H.M. Cheng, Chem. Commun. 50 (74)
(2014) 10837-10840.
[86] Y. Wang, X. Bai, H. Qin, F. Wang, Y. Li, X. Li, S. Kang, Y. Zuo L. Cui, ACS
Appl. Mater. Interfaces. 8 (27) (2016) 17212-17219.
[87] Z. Li, Y. Zhou, J. Zhang, W. Tu, Q. Liu, T. Yu Z. Zou, Cryst. Growth Des.
12 (3) (2012) 1476-1481.
[88] J. Sun, J. Zhang, M. Zhang, M. Antonietti, X. Fu X. Wang, Nat. Commun. 3
(2012) 1139.
[89] P. Yaashikaa, P.S. Kumar, S.J. Varjani A. Saravanan, J. CO2 Utili. 33(2019)
131-147.
[90] M.A. ScibiohB. Viswanathan, Carbon Dioxide to chemicals and fuels,
Elsevier, 2018.
[91] N. Sutin, C. Creutz E. Fujita, Comment. Inorg. Chem. 19 (2) (1997) 67-92.
[92] Q. Xiang, B. Cheng J. Yu, Angew. Chem. Int. Ed. 54 (39) (2015) 11350-
11366.
[93] V.P. Indrakanti, J.D. Kubicki H.H. Schobert, Energy Env. Sci. 2 (7) (2009)
745-758.
[94] M.A. Méndez, P. Voyame H.H. Girault, Angew. Chem. Int. Ed. 50 (32) (2011)
7391-7394.
[95] C.W. Machan, S.A. Chabolla, J. Yin, M.K. Gilson, F.A. Tezcan C.P. Kubiak, J.
Am. Chem. Soc. 136 (41) (2014) 14598-14607.
[96] V.S. Thoi, N. Kornienko, C.G. Margarit, P. Yang C.J. Chang, J. Am. Chem.
Soc. 135 (38) (2013) 14413-14424.
[97] D. Hong, Y. Tsukakoshi, H. Kotani, T. Ishizuka T. Kojima, J. Am. Chem. Soc.
139 (19) (2017) 6538-6541.
[98] H. Ishida, T. Terada, K. Tanaka T. Tanaka, Inorg. Chem. 29 (5) (1990) 905-
911.
[99] Y. Tamaki, T. Morimoto, K. Koike O. Ishitani, Proceed. Nat. Acad. Sci. 109
(39) (2012) 15673-15678.
[100] M.D. Sampson, A.D. Nguyen, K.A. Grice, C.E. Moore, A.L. Rheingold C.P.
Kubiak, J. Am. Chem. Soc. 136 (14) (2014) 5460-5471.
[101] J. Chauvin, F. Lafolet, S. Chardon‐Noblat, A. Deronzier, M. Jakonen M.
Haukka, Chem.: Eu. J. 17 (15) (2011) 4313-4322.
[102] C.M. Bolinger, N. Story, B.P. Sullivan T.J. Meyer, Inorg. Chem. 27 (25)
(1988) 4582-4587.
[103] J. Hawecker, J.-M. Lehn R. Ziessel, J. Chem. Soc., Chemi. Commun. (9)
(1983) 536-538.
[104] M.N. Collomb-Dunand-Sauthier, A. Deronzier R. Ziessel, J. Organomet.
Chem. 444 (1-2) (1993) 191-198.
[105] J.M. Kelly, C.M. O'Connell J.G. Vos, J. Chem. Soc., Dalton. Trans. (2) (1986)
253-258.
[106] R. Ziessel, J. Hawecker, J.M. Lehn, Helvetica Chim. Acta. 69 (5) (1986)
1065-1084.
[107] R.O. Reithmeier, S. Meister, B. Rieger, A. Siebel, M. Tschurl, U. Heiz E.
Herdtweck, Dalton. Trans. 43 (35) (2014) 13259-13269.
[108] K. Mochizuki, S. Manaka, I. Takeda T. Kondo, Inorg. Chem. 35 (18) (1996)
5132-5136.
[109] A.A. Khan, M. Tahir, J. CO2 Utili. 29) (2019) 205-239.
[110] Y. Kuramochi, O. Ishitani, Inorg. Chem. 55 (11) (2016) 5702-5709.
[111] C. Costentin, M. Robert J.-M. Savéant, Chem. Soc. Rev. 42 (6) (2013) 2423-
2436.
[112] J.M. Savéant, Chem. Rev. 108 (7) (2008) 2348-2378.
[113] B.S. Kwak, K. Vignesh, N.K. Park, H.J. Ryu, J.I. Baek M.J.F. Kang, Fuel.
143 (2015) 570-576.
[114] I.A. Shkrob, T.W. Marin, H. He, P. Zapol, J. Phys. Chem. C. 116 (17) (2012)
9450-9460.
[115] W.N. Wang, J. Soulis, Y.J. Yang P. Biswas. Aerosol Air Qua. Res. 14 (2)
(2014) 533-549.
[116] G. Liu, L. Wang, H.G. Yang, H.-M. Cheng G.Q. Lu, J. Mater. Chem. 20 (5)
(2010) 831-843.
[117] A. Bachmeier, F. Armstrong. Curr. Opin. Chem. Biol. 25 (2015) 141-151.
[118] J.G.M. Winkelman, O. Voorwinde, M. Ottens, A. Beenackers L. Janssen,
Chem. Eng. Sci. 57 (19) (2002) 4067-4076.
[119] I. Ganesh, Mater. Sci. App. 2 (10) (2011) 1407.
[120] M. Khalil, J. Gunlazuardi, T.A. Ivandini, A. Umar, Renew. Sust. Energy Rev.
113 (2019) 109246.
[121] G. Sahara, O. Ishitani, Inorg. Chem. 54 (11) (2015) 5096-5104.
[122] Y. Yamazaki, H. Takeda, O. Ishitani J. Photoch. Photobio. C. 25 (2015) 106-
137.
[123] G. Kyriakou, M.B. Boucher, A.D. Jewell, E.A. Lewis, T.J. Lawton, A.E.
Baber, H.L. Tierney, M. Flytzani-Stephanopoulos E.C.H. Sykes, Sci. 335
(6073) (2012) 1209-1212.
[124] W. Liu, L. Zhang, W. Yan, X. Liu, X. Yang, S. Miao, W. Wang, A. Wang T.
Zhang, Chem. Sci. 7 (9) (2016) 5758-5764.
[125] A. Wang, J. Li, T. Zhang, Nat. Rev. Chem. 2 (6) (2018) 65-81.
[126] A. Nakada, K. Koike, K. Maeda O. Ishitani, Green Chem. 18 (1) (2016) 139-
143.
[127] B. Kumar, M. Llorente, J. Froehlich, T. Dang, A. Sathrum C.P. Kubiak, Annu.
Rev. Phys. Chem. 63 (2012) 541-569.
[128] Q. Yi, W. Li, J. Feng K. Xie, Chem. Soc. Rev. 44 (15) (2015) 5409-5445.
[129] J. Shi, Y. Jiang, Z. Jiang, X. Wang, X. Wang, S. Zhang, P. Han C. Yang,
Chem. Soc. Rev. 44 (17) (2015) 5981-6000.
[130] T.W. Woolerton, S. Sheard, E. Reisner, E. Pierce, S.W. Ragsdale F.A.
Armstrong, J. Am. Chem. Soc. 132 (7) (2010) 2132-2133.
[131] A. Parkin, J. Seravalli, K.A. Vincent, S.W. Ragsdale F.A. Armstrong, J. Am.
Chem. Soc. 129 (34) (2007) 10328-10329.
[132] C.K. SavileJ.J. Lalonde, Curr. Opin. Biotech. 22 (6) (2011) 818-823.
[133] I.M. MachadoS. Atsumi, J. Biotech. 162 (1) (2012) 50-56.
[134] E. Jacob-Lopes, C.H.G. Scoparo, L.M.C.F. Lacerda T.T. Franco, Chem. Eng.
Process: Process Intensification. 48 (1) (2009) 306-310.
[135] A. Scragg, A. Illman, A. Carden S. Shales, Biomass Bioenerg. 23 (1) (2002)
67-73.
[136] M.G. de MoraisJ.A.V. Costa, Energy Convers. Manage. 48 (7) (2007) 2169-
2173.
[137] H.J. Ryu, K.K. Oh Y.S. Kim, J. Ind. Eng. Che. 15 (4) (2009) 471-475.
[138] S.Y. Chiu, M.T. Tsai, C.Y. Kao, S.C. Ong C.S. Lin, Eng. Life Sci. 9 (3) (2009)
254-260.
[139] L. Cheng, L. Zhang, H. Chen C. Gao, Sep. Purif. Tech. 50 (3) (2006) 324-
329.
[140] L.-H. Fan, Y.-T. Zhang, L. Zhang H.-L. Chen, J. Membrane Sci. 325 (1)
(2008) 336-345.
[141] N. Sakai, Y. Sakamoto, N. Kishimoto, M. Chihara I. Karube, Energy Convers.
Manage. 36 (6-9) (1995) 693-696.
[142] S. Chae, E. Hwang H.-S. Shin, Bioresource Tech. 97 (2) (2006) 322-329.
[143] H.-F. Jin, B.-R. Lim K. Lee, J. Environ. Sci. Heal. A. 41 (12) (2006) 2813-
2824.
[144] M. Negoro, N. Shioji, K. Miyamoto Y. Micira, App. Biochem. Biotech. 28 (1)
(1991) 877.
[145] M.G. De Morais, J.A.V. Costa, Biotech. Lett. 29 (9) (2007) 1349-1352.
[146] C. Yoo, S.-Y. Jun, J.-Y. Lee, C.-Y. Ahn H.-M. Oh, Bioresource Tech. 101 (1)
(2010) S71-S74.
[147] S.-H. Ho, W.-M. Chen J.-S. Chang, Bioresource Tech. 101 (22) (2010) 8725-
8730.
[148] M.G. De MoraisJ.A.V. Costa, J. Biotech. 129 (3) (2007) 439-445.
[149] A. Melis, Energy Env. Sci. 5 (2) (2012) 5531-5539.
[150] A.S. Agarwal, Y. Zhai, D. Hill N. Sridhar, Chem. Sus. Chem. 4 (9) (2011)
1301-1310.
[151] W. Tang, A.A. Peterson, A.S. Varela, Z.P. Jovanov, L. Bech, W.J. Durand, S.
Dahl, J.K. Nørskov I. Chorkendorff, Phy. Chem. Chem. Phy. 14 (1) (2012)
76-81.
[152] E. Barton Cole, P.S. Lakkaraju, D.M. Rampulla, A.J. Morris, E. Abelev A.B.
Bocarsly, J. Am. Chem. Soc. 132 (33) (2010) 11539-11551.
[153] S. Ma, P.J. Kenis, Curr. Opin. Chem. Eng. 2 (2) (2013) 191-199.
[154] A.A. Peterson, F. Abild-Pedersen, F. Studt, J. Rossmeisl J.K. Nørskov, Energy
Env. Sci. 3 (9) (2010) 1311-1315.
[155] J. Wu, F.G. Risalvato, F.-S. Ke, P. Pellechia X.-D. Zhou, J. Electrochem. Soc.
159 (7) (2012) F353-F359.
[156] H.-R.M. Jhong, F.R. Brushett, L. Yin, D.M. Stevenson P.J. Kenis, J.
Electrochem. Soc. 159 (3) (2012) B292-B298.
[157] B.C. Marepally, C. Ampelli, C. Genovese, T. Saboo, S. Perathoner, F.M.
Wisser, L. Veyre, J. Canivet, E.A. Quadrelli G. Centi, Chem. Sus. Chem. 10
(22) (2017) 4442-4446.
[158] C. Ampelli, C. Genovese, M. Errahali, G. Gatti, L. Marchese, S. Perathoner G.
Centi, J. App. Electrochem. 45 (7) (2015) 701-713.
[159] S. Prince-Richard, M. Whale N. Djilali, Int. J. Hydrogen Energy. 30 (11)
(2005) 1159-1179.
[160] S. Kaneco, R. Iwao, K. Iiba, K. Ohta T. Mizuno, Energy. 23 (12) (1998)
1107-1112.
[161] C. Ampelli, C. Genovese, G. Centi, R. Passalacqua S. Perathoner, Top. Catal.
59 (8-9) (2016) 757-771.
[162] F. Urbain, P. Tang, N.M. Carretero, T. Andreu, L.G. Gerling, C. Voz, J.
Arbiol J.R. Morante, Energy Env. Sci. 10 (10) (2017) 2256-2266.
[163] Y. Oh, H. Vrubel, S. Guidoux X. Hu, Chem. Commun. 50 (29) (2014) 3878-
3881.
[164] R. ChaplinA. Wragg, J. App. Electrochem. 33 (12) (2003) 1107-1123.
[165] M. Azuma, K. Hashimoto, M. Hiramoto, M. Watanabe T. Sakata, J.
Electrochem. Soc. 137 (6) (1990) 1772-1778.
[166] B.A. Rosen, A. Salehi-Khojin, M.R. Thorson, W. Zhu, D.T. Whipple, P.J.
Kenis R.I. Masel, Science. 334 (6056) (2011) 643-644.
[167] J. Lee, J. Lim, C.-W. Roh, H.S. Whang H. Lee, J. CO2 Util. 31 (2019) 244-
250.
[168] T. Hatsukade, K.P. Kuhl, E.R. Cave, D.N. Abram T.F. Jaramillo, Phy. Chem.
Chem. Phy. 16 (27) (2014) 13814-13819.
[169] S.Y. Lee, H. Jung, N.-K. Kim, H.-S. Oh, B.K. Min Y.J. Hwang, J. Am. Chem.
Soc. 140 (28) (2018) 8681-8689.
[170] E.L. ClarkA.T. Bell, J. Am. Chem. Soc. 140 (22) (2018) 7012-7020.
[171] S. Ren, D. Joulié, D. Salvatore, K. Torbensen, M. Wang, M. Robert C.P.
Berlinguette, Science. 365 (6451) (2019) 367-369.
[172] C.-T. Dinh, T. Burdyny, M.G. Kibria, A. Seifitokaldani, C.M. Gabardo, F.P.G.
de Arquer, A. Kiani, J.P. Edwards, P. De Luna O.S. Bushuyev, Science. 360
(6390) (2018) 783-787.
[173] J. Albo, G. Beobide, P. Castaño A. Irabien, J. CO2 Util. 18 (2017) 164-172.
[174] S. Narayanan, B. Haines, J. Soler T. Valdez, J. Electrochem. Soc. 158 (2)
(2011) A167-A173.
[175] R. Kas, R. Kortlever, H. Yılmaz, M.T. Koper G. Mul, Chem. Electro. Chem.
2 (3) (2015) 354-358.
[176] D.T. Whipple, E.C. Finke P.J. Kenis, Electrochem. Solid-State Lett. 13 (9)
(2010) 109-111.
[177] S. Ma, M. Sadakiyo, R. Luo, M. Heima, M. Yamauchi P.J. Kenis, J. Power
Sources. 301) (2016) 219-228.
[178] K. Jiang, R.B. Sandberg, A.J. Akey, X. Liu, D.C. Bell, J.K. Nørskov, K. Chan
H. Wang, Nat. Catal. 1 (2) (2018) 111.
[179] F. Pan, B. Li, W. Deng, Z. Du, Y. Gang, G. Wang Y. Li, App. Catal. B: Env.
252) (2019) 240-249.
[180] S.D. Ebbesen, R. Knibbe M. Mogensen, J. Electrochem. Soc. 159 (8) (2012)
F482-F489.
[181] A. Fridman, Plasma chemistry, Cambridge university press, 2008.
[182] M.U. Khan, Y. Zhao, T. Hui, M.K. Shahzad, H. Cui, D. Zhao, Opt. Express.
27 (12) (2019) 16738-16750.
[183] A. Bogaerts, T. Kozák, K. Van Laer R. Snoeckx, Faraday Discuss. 183 (2015)
217-232.
[184] R. Asisov, V. Givotov, E. Krasheninnikov, B. Potapkin, V. Rusanov A.
Fridman, in Proc. 5th Intern. Symp. Plasma Chemistry. 1981 1981.
[185] X. Tao, M. Bai, X. Li, H. Long, S. Shang, Y. Yin X. Dai, Prog. Energy
Combust. Sci. 37 (2) (2011) 113-124.
[186] H. SekiguchiY. Mori, Thin Solid Films. 435 (1-2) (2003) 44-48.
[187] A. Fridman, A. Chirokov A. Gutsol, J. Phy. D: App. Phys. 38 (2) (2005).
[188] M. Moreau, N. Orange M. Feuilloley, Biotech. Advan. 26 (6) (2008) 610-617.
[189] Z. Bo, J. Yan, X. Li, Y. Chi K. Cen, Int. J. Hydrogen Energy. 33 (20) (2008)
5545-5553.
[190] U. Kogelschatz, Plasma Chem. Plasma P. 23 (1) (2003) 1-46.
[191] U. Kogelschatz, B. Eliasson W. Egli, Pure App. Chem. 71 (10) (1999) 1819-
1828.
[192] Istadi, N.A.S. Amin, Fuel. 85 (5-6) (2006) 577-592.
[193] X. Duten, D. Packan, L. Yu, C. Laux C. Kruger, IEEE T. Plasma Sci. 30 (1)
(2002) 178-179.
[194] A.-A. Mohamed, R. Block, K.H. Schoenbach, IEEE T. Plasma Sci. 30 (1)
(2002) 182-183.
[195] X. Li, X. Tao Y. Yin, IEEE T. Plasma Sci. 37 (6) (2009) 759-763.
[196] B. Eliasson, U. Kogelschatz, B. Xue L.-M. Zhou, Ind. Eng. Chem. Res. 37 (8)
(1998) 3350-3357.
[197] E. Jwa, S. Lee, H. Lee Y. Mok, Fuel Process. Technol. 108) (2013) 89-93.
[198] M. Nizio, A. Albarazi, S. Cavadias, J. Amouroux, M.E. Galvez P. Da Costa,
Int. J. Hydrogen Energy. 41 (27) (2016) 11584-11592.
[199] Y. ZengX. Tu, IEEE T. Plasma Sci. 44 (4) (2015) 405-411.
[200] C. De Bie, J. van Dijk A. Bogaerts, J. Phys. Chem. C. 120 (44) (2016)
25210-25224.
[201] L. Maya, J. Vacuum Sci. Tech. A: Vacuum, Surfaces, and Films. 18 (1) (2000)
285-287.
[202] M. Kano, G. Satoh S. Iizuka, Plasma Chem. Plasma Process. 32 (2) (2012)
177-185.
[203] F. Javier, S.H. Moreno, A.I. Stankiewicz G.D. Stefanidis, Int. J. Hydrogen
Energy. 42 (18) (2017) 12943-12955.
[204] G. Chen, N. Britun, T. Godfroid, V. Georgieva, R. Snyders M.-P. Delplancke-
Ogletree, J. Phys. D: App. Phy. 50 (8) (2017) 084001.
[205] N. Hayashi, T. Yamakawa S. Baba, Vacuum. 80 (11-12) (2006) 1299-1304.
[206] J.-Q. Zhang, Y.-J. Yang, J.-S. Zhang Q. Liu, Acta Chim. Sinica-Chinese Ed.
60 (11) (2002) 1973-1980.
[207] M.-w. Li, G.-h. Xu, Y.-l. Tian, L. Chen H.-f. Fu, J. Phy. Chem. A. 108 (10)
(2004) 1687-1693.
[208] Y. Yang, Ind. Eng. Chem. Res. 41 (24) (2002) 5918-5926.
[209] A. Indarto, J.-W. Choi, H. Lee H.K. Song, Energy. 31 (14) (2006) 2986-2995.
[210] H.K. Song, J.-W. Choi, S.H. Yue, H. Lee B.-K. Na, Catal. Today. 89 (1-2)
(2004) 27-33.
[211] Q. Wang, B.-H. Yan, Y. Jin Y. Cheng, Plasma Chem. Plasma Process. 29 (3)
(2009) 217-228.
[212] H. Long, S. Shang, X. Tao, Y. Yin X. Dai, Int. J. Hydrogen Energy. 33 (20)
(2008) 5510-5515.
[213] X. Tao, M. Bai, Q. Wu, Z. Huang, Y. Yin X. Dai, Int. J. Hydrogen Energy. 34
(23) (2009) 9373-9378.
[214] T.-s. Lan, W. Ran, H.-l. Long, Y.-j. WANG Y.-x. YIN, Natural Gas Industry.
27 (5) (2007) 129.
[215] X. Tao, F. Qi, Y. Yin X. Dai, Int. J. Hydrogen Energy. 33 (4) (2008) 1262-
1265.
[216] B.G. Lougou, J. Hong, Y. Shuai, X. Huang, Y. Yuan H. Tan, Sol. Energy 148)
(2017) 117-127.
[217] B. Guene Lougou, Y. Shuai, G. Chaffa, C. Ahouannou, R. Pan, H. Zhang, H.
Tan, Energy Technology. 7 (3) (2019) 1800588.
[218] M. Romero, A.J.E. Steinfeld E. Science. 5 (11) (2012) 9234-9245.
[219] B.G. Lougou, Y. Shuai, R. Pan, G. Chaffa, H.J.I.J.o.H. Tan M. Transfer. 127)
(2018) 61-74.
[220] B.G. Lougou, Y. Shuai, H. Xing, Y. Yuan, H. Tan, Int. J. Heat Mass Tran.
111 (2017) 410-418.
[221] C. Agrafiotis, H. von Storch, M. Roeb C. Sattler, Renew. Sust. Energy Rev.
29 (2014) 656-682.
[222] M. Böhmer, U. Langnickel M. Sanchez, Sol. Energy Mater. 24 (1-4) (1991)
441-448.
[223] I. Spiewak, C.E. Tyner U. Langnickel, Applications of solar reforming
technology, Sandia National Labs., Albuquerque, NM (United States), 1993.
[224] N. Gokon, S. Nakamura, T. Hatamachi T. Kodama, Energy. 68) (2014) 773-
782.
[225] A. Giaconia, G. Monteleone, B. Morico, A. Salladini, K. Shabtai, M.
Sheintuch, D. Boettge, J. Adler, V. Palma S. Voutetakis, Energy Procedia. 69)
(2015) 1750-1758.
[226] J.K. Dahl, A.W. Weimer, A. Lewandowski, C. Bingham, F. Bruetsch A.
Steinfeld, Ind. Eng. Chem. Res. 43 (18) (2004) 5489-5495.
[227] M. Epstein, I. Spiewak, A. Segal, I. Levy, D. Lieberman, M. Meri V. Lerner,
in Eighth Int. Symposium on Sol. Therm. Concentrating Tech. 1996.
[228] A. WörnerR. Tamme, Catal. Today. 46 (2-3) (1998) 165-174.
[229] S. Möller, SolarPACES Annual Report. 2007) (2008) 4.5-4.6.
[230] J. Petrasch, B. Schrader, P. Wyss A. Steinfeld, J. Heat Transfer. 130 (3) (2008)
032602.
[231] J. PetraschA. Steinfeld, Chem. Eng. Sci. 62 (16) (2007) 4214-4228.
[232] T. Kodama, N. Gokon, K. Matsubara, K. Yoshida, S. Koikari, Y. Nagase K.
Nakamura, Energy Procedia. 49) (2014) 1990-1998.
[233] J. Hinkley.) (2013).
[234] J.F. Muir, R.E. Hogan Jr, R.D. Skocypec R. Buck, Sol. Energy. 52 (6) (1994)
467-477.
[235] R. Buck, J.F. Muir R.E. Hogan, Sol. Energy Mater. 24 (1-4) (1991) 449-463.
[236] A. Berman, R.K. Karn M. Epstein, Energy & Fuels. 20 (2) (2006) 455-462.
[237] R. RubinJ. Karni, J. Sol. Energy Eng. 133 (2) (2011) 021008.
[238] C. Sugarmen, A. Rotstein, U. Fisher J. Sinai, J. Sol. Energy Eng. 126 (3)
(2004) 867-871.
[239] S. Möller, D. Kaucic C. Sattler, J. Sol. Energy Eng. 128 (1) (2006) 16-23.
[240] R. Tamme, R. Buck, M. Epstein, U. Fisher C. Sugarmen, J. Sol. Energy Eng.
123 (2) (2001) 160-163.
[241] R. Diver, J. Fish, R. Levitan, M. Levy, E. Meirovitch, H. Rosin, S.
Paripatyadar J. Richardson, Sol. Energy. 48 (1) (1992) 21-30.
[242] S. ParipatyadarJ. Richardson, Sol. energy. 41 (5) (1988) 475-485.
[243] J.R. Scheffe, A. Steinfeld, Mater. Today. 17 (7) (2014) 341-348.
[244] J.R. Scheffe, D. Weibel, A.J.E. Steinfeld, Fuels. 27 (8) (2013) 4250-4257.
[245] A.H. McDaniel, E.C. Miller, D. Arifin, A. Ambrosini, E.N. Coker, R. O'Hayre,
W.C. Chueh, J. Tong, Energy Env. Sci. 6 (8) (2013) 2424-2428.
[246] A. Evdou, V. Zaspalis L. Nalbandian, Fuel. 89 (6) (2010) 1265-1273.
[247] D. Sastre, A.J. Carrillo, D.P. Serrano, P. Pizarro J. Coronado, Top. Catal. 60
(15-16) (2017) 1108-1118.
[248] A. Demont, S. Abanades, RSC. Adv. 4 (97) (2014) 54885-54891.
[249] A.H. McDaniel, A. Ambrosini, E.N. Coker, J.E. Miller, W.C. Chueh, R.
O’Hayre J. Tong, Energy Procedia. 49 (2014) 2009-2018.
[250] A. Demont, S.p. Abanades, E. Beche, J. Phys. Chem. C. 118 (24) (2014)
12682-12692.
[251] M. Ezbiri, M. Takacs, B. Stolz, J. Lungthok, A. Steinfeld R. Michalsky, J.
Mater. Chem. A. 5 (29) (2017) 15105-15115.
[252] C.L. Muhich, S. Blaser, M.C. Hoes, A. Steinfeld, Int. J. Hydrogen Energy. 43
(41) (2018) 18814-18831.
[253] H. Kong, Y. Hao, H. Jin, App. Energy. 228 (2018) 301-308.
[254] Y. Lu, L. Zhu, C. Agrafiotis, J. Vieten, M. Roeb, C. Sattler, Prog. Energy
Combust. C. Science. 75) (2019) 100785.
[255] S. Abanades, G. Flamant, Sol. Energy. 80 (12) (2006) 1611-1623.
[256] W.C. Chueh, S.M. Haile, Chem. Sus. Chem. 2 (8) (2009) 735-739.
[257] W.C. Chueh, C. Falter, M. Abbott, D. Scipio, P. Furler, S.M. Haile, A.
Steinfeld, Science. 330 (6012) (2010) 1797-1801.
[258] R. Panlener, R. Blumenthal, J.E. Garnier, J. Phy. Chem. Solids. 36 (11)
(1975) 1213-1222.
[259] O.T. Sørensen, J. Solid State Chem. 18 (3) (1976) 217-233.
[260] P. Furler, J.R. Scheffe, A. Steinfeld, Energy Env. Sci. 5 (3) (2012) 6098-
6103.
[261] M.A. Panhans, R.N. Blumenthal, Solid State Ion. 60 (4) (1993) 279-298.
[262] A.S. Jonathan, R. Scheffe, Mater. Today. 17 (7) (2014) 341-348.
[263] E.N. Coker, J.A. Ohlhausen, A. Ambrosini J.E. Miller, J. Mater. Chem. 22
(14) (2012) 6726-6732.
[264] W.C. Chueh, S.M. Haile, Philos. T. R. Soc. A. 368 (1923) (2010) 3269-3294.
[265] G.J. VanHandel, R.N. Blumenthal, J. Electrochem. Soc. 121 (9) (1974) 1198-
1202.
[266] J.E. Miller, M.D. Allendorf, R.B. Diver, L.R. Evans, N.P. Siegel J.N. Stuecker,
J. Mater. Sci. 43 (14) (2008) 4714-4728.
[267] H. Aoki, H. Kaneko, N. Hasegawa, H. Ishihara, A. Suzuki, Y. Tamaura, Solid
State Ion. 172 (1-4) (2004) 113-116.
[268] Y. Tamaura, Y. Ueda, J. Matsunami, N. Hasegawa, M. Nezuka, T. Sano M.
Tsuji, Sol. Energy. 65 (1) (1999) 55-57.
[269] C. Agrafiotis, A. Zygogianni, C. Pagkoura, M. Kostoglou A.G.
Konstandopoulos, AIChE J. 59 (4) (2013) 1213-1225.
[270] M. Inoue, N. Hasegawa, R. Uehara, N. Gokon, H. Kaneko Y. Tamaura, Sol.
Energy. 76 (1-3) (2004) 309-315.
[271] Y. Tamaura, M. Kojima, T. Sano, Y. Ueda, N. Hasegawa, M. Tsuji, Int. J.
Hydrogen Energy. 23 (12) (1998) 1185-1191.
[272] G.L. Bachirou, Y. Shuai, J. Zhang, X. Huang, Y. Yuan H. Tan, Int. J.
Hydrogen Energy. 41 (44) (2016) 19936-19946.
[273] C. Agrafiotis, M. Roeb, C. Sattler, Renew. Sust. Energy Rev. 42 (2015) 254-
285.
[274] M.D. Allendorf, R.B. Diver, N.P. Siegel, J.E. Miller, Energy and Fuels. 22 (6)
(2008) 4115-4124.
[275] C.C. Agrafiotis, C. Pagkoura, A. Zygogianni, G. Karagiannakis, M. Kostoglou
A.G. Konstandopoulos, Int. J. Hydrogen Energy. 37 (11) (2012) 8964-8980.
[276] M. Kostoglou, S. Lorentzou A.G. Konstandopoulos, Int. J. Hydrogen Energy.
39 (12) (2014) 6317-6327.
[277] M. Neises, M. Roeb, M. Schmücker, C. Sattler, R. Pitz‐Paal, Int. Energy Res.
34 (8) (2010) 651-661.
[278] K.S. Go, S.R. Son S. Kim, Int. J. Hydrogen Energy. 33 (21) (2008) 5986-
5995.
TOC
The continuous burning of fossil fuels to meet the energy demands results in
excessive emissions of CO2 in the atmosphere which consequently result in climate
change. The utilization of CO2 into solar fuels can help mitigate the consumption of
fossil fuel and excessive emissions of CO2. This review discusses the latest
developments in those technologies that are used to convert CO2 into solar fuels and
useful chemicals.

You might also like