0% found this document useful (0 votes)
146 views56 pages

Progress in Energy and Combustion Science: Turgut M. Gür

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
146 views56 pages

Progress in Energy and Combustion Science: Turgut M. Gür

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 56

Progress in Energy and Combustion Science 89 (2022) 100965

Contents lists available at ScienceDirect

Progress in Energy and Combustion Science


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

Carbon Dioxide Emissions, Capture, Storage and Utilization: Review of


Materials, Processes and Technologies
Turgut M. Gür
Department of Materials Science and Engineering, Stanford University, Stanford, CA 94305, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Carbon capture and storage (CCS) is an essential component of mitigating climate change, which arguably
Carbon dioxide presents an existential challenge to our planet. Although CO2 emissions have been on the global agenda for
Carbon capture and storage several decades, progress has been extremely slow, insufficient and sporadic.
CO2 emissions
Anthropogenic CO2 emissions are the direct result of our addiction to fossil fuels, and in 2018 accounted for
Pre-combustion
Post-combustion
68% (or, 37.5 GtCO2) out of the total greenhouse gas (GHG) emissions of 55.3 GtCO2e globally. Capturing CO2 at
Oxy-combustion such massive quantities would require resources and technologies that can operate cost-effectively at the multi-
Direct air capture gigaton scale, which we currently lack. Moreover, CO2 capture is an expensive and highly energy intensive
CO2 utilization process complicated further by logistics and the diversity of the emission sources that vary by volume,
CO2 sorbent materials composition, location, type, and industry or sector. At the same time, however, such diversity also highlights that
Oxygen-based conversion one size does not fit all, and hence, dictates the need for a multi-prong strategy that emphasizes the necessity to
Climate change develop wide range of CCS technologies, materials and processes.
This article presents a global overview and impartial assessment of the current state of CCS challenges in an
extensive manner covered under the main headings of pre- and post-combustion CO2 capture, direct air capture,
CO2 transport and storage and utilization, and carbon pricing. Materials aspects of post-combustion CO2 capture
technologies are reviewed in detail. The article provides critical discussions of fundamental phenomena and
recent advances in the field, as well as tutorial-type background information, where appropriate. The article
reviews the status of global CO2 emissions as well as carbon sources and sinks, and examines a broad range of
major technologies, methodologies, processes, and materials for CO2 capture, discusses technology options for
carbon capture from fossil fuel-based power generation, presents the challenges to storage, utilization and the
global pricing of CO2, and finishes with an assessment of knowledge gaps, mitigation options and opportunities
for advances.
The article emphasizes the fact that there are no easy fixes or cheap technological solutions to the inter­
connected problems of energy, CO2 emissions and climate change. The threats to Earth’s ecosystems are too real
and imminent to be judged and driven only by economics. In this existential context, the choice between ‘pay
now’ or ‘pay later’ is clear and paying later will be much more expensive. The world must act now.

1. Introduction and their continued use is expected to increase by 2% annually through


2030, which would double the required amount of reductions to limit
It was not an understatement when in December 2018 the Secretary the temperature rise to 1.5 ◦ C [4].
General Antonio Guterres of the United Nations (UN) observed, The power generation sector driven primarily by fossil fuels
“We are in trouble. We are in deep trouble with climate change…” [1] during accounted for 36% of the CO2 emissions across advanced economies of
the opening of the 24th annual UN Climate Change Conference held in the world in 2019, down from 42% in 2012 [6]. Although coal’s share in
Poland [2]. In early December 2020, he voiced his warning bell again primary energy fell to a 15-year low of 27.2%, it still remains a signif­
and called out “Humanity is waging war on nature. This is suicidal.” [3]. icant source for the global energy mix. Coal consumption in the US,
Indeed, the Production Gap Report 2020 indicates that despite the India and China accounted for, respectively, 8.4%, 12.0% and 50.5% of
impending climate crisis, the world insists on investing in fossil fuels, the global coal consumption in 2018 [7]. While China was the biggest

E-mail address: [email protected].

https://doi.org/10.1016/j.pecs.2021.100965
Received 2 February 2021; Received in revised form 12 September 2021; Accepted 12 September 2021
Available online 18 December 2021
0360-1285/© 2021 Elsevier Ltd. All rights reserved.
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

contributor to the global CO2 emissions in 2018 at 27.8%, US came in Unfortunately, the global energy storage capacity at the end of 2020
2nd place with 15.2% [8]. stood at merely 800 GWh [16] and 95% of this storage capacity is
Fossil fuels have collectively been the single largest source provided by pumped hydro alone, which is largely inaccessible to the
of CO2 emissions, responsible for nearly 65% of the global green wider global community. By the end of 2019, the world had installed a
house gas emissions [9]. As global emissions increased by nearly 1.5% total of 2533 GW of renewable capacity [17]. To store 2533 GW of
(or, 450 Mt CO2) in 2017, developing countries in Asia were largely renewable power for an average duration of four hours, approximately
responsible for this increase, which is led primarily by the largest CO2 10,000 GWh of storage capability would be required. It is clear that 800
emitter China, followed by India. This is illustrated in Fig. 1, which GWh of storage capacity falls far too short of meeting the need by any
provides the historical record of CO2 emissions by major emitters be­ measure. Consequently, without synchronous build up of commensurate
tween 1990 and 2018. storage capacity, the rate of growth and penetration of renewable power
Recent assessment from the Intergovernmental Panel on Climate sources into the electric grid to dominate the global energy economy
Change (IPCC) recommends limiting the cumulative quantity of CO2 may run the risk of slowing down.
emissions between 2018 and the start of achieving net-zero global Besides, expectations that renewables are able to provide 100% of
emissions (i.e., the world’s remaining carbon budget) to 750 GtCO2 for the global energy needs may also be unrealistic, and even be impractical.
an even chance of restraining global warming to 1.5 ◦ C of temperature Whether the world’s entire energy needs can truly be supplied by re­
rise, or to 550 GtCO2 for a 2/3rd chance [10]. This would imply reaching newables only [18,19] is an ongoing debate [20,21] with arguments on
carbon neutrality in 35 years for the 750 GtCO2 budget and nearly both sides of the issue. It’s not clear how fast the world can quickly shake
25 years for the 550 GtCO2 budget scenarios both of which imminently off its addiction to fossil fuels, and instead fully embrace renewables as
demand urgent action on the part of the global community at an un­ its only energy source.
precedented scale. Energy and climate cannot be considered independent of each other.
The 2019 UN Emissions Gap Report [11] and the recent IPCC reports They are intimately linked as CO2 emissions impact climate change. So
[12,13] strongly urged a 55% cut in CO2 emissions by 2030 in order to solutions to mitigate climate change must be commensurate with and
limit global warming to within 1.5 ◦ C. The reports also urged rapid responsive to the rising energy demands of the global community.
transformation of the global energy systems from fossil fuels to renew­ Energy is also intimately linked with environment, food production, and
ables in order to achieve net-zero emissions by mid-century. fresh water availability, all of which also affect population shifts,
The world consumed 654 EJ of energy in 2018, of which 80% was growth, and mass migrations. For example, electricity is an essential
supplied by fossil fuels that are also expected to provide 68% of the component for both the extraction and distribution of water for human
global energy in 2050 [14]. By comparison, the global community activity and needs. At the same time, water is the lifeblood of
consumed 80.3 EJ (or, 22,315 TWh) of electricity in 2018. For 100% thermoelectric power production. Understanding such complex
renewables, this would require building a renewable electricity gener­ interlinks among these individual elements, i.e., the “grand nexus” [22,
ation capacity of 181,700 TWh, which is more than 8 times the total 23], is essential for proper assessment of energy options, especially
installed generation capacity we now have. Transforming the global vis-à-vis carbon capture. The strong interlinks and dependencies among
energy infrastructure into solely renewables will be an unprecedented the major components of the grand nexus are illustrated in Fig. 2. Water
undertaking, especially when considering the current generation ca­ is a critically scarce resource in many parts of the world. It is also
pacity is only 5490 TWh for all renewables. With the exception of indispensible for thermoelectric power generation, which requires close
pumped hydro and compressed air energy storage, both of which are to 1 m3/kWh for natural gas combined cycle plants, while coal base
site-specific and available in only certain geographies, other renewable combustion plants demand >2 m3/kWh as depicted in Fig. 3. Moreover,
sources are undispatchable in nature and have inherently variable and CO2 capture processes place even greater demand for water as high­
intermittent power outputs, which make it difficult to integrate into the lighted in Fig. 3, where the water demand nearly doubles with carbon
global electric grid without adequate energy storage capacity [15]. capture from various fossil fuel-based electricity production technolo­
gies [24]. Accordingly, any technology advancement in electric power
generation from fossil fuels that helps reduce the demand for water
presents huge environmental and economic benefits. In that regard,
oxygen-based technologies [22] discussed in the present article largely

Fig. 2. Schematic depiction of the ‘grand nexus’ highlighting the strong in­
terlinks and interdependencies among its major components. Reprinted from
[22], with permission from Elsevier.
Fig. 1. Top greenhouse gas emitting countries, highlighting China’s substantial
contribution to growth of emissions (in gigaton of CO2 equivalents: GtCO2e) [5].

2
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 3. Water withdrawal rates for thermoelectric power generation from coal and natural gas with and without carbon capture, illustrating up to ~100% increase in
water demand for capture. CC: Combined cycle for natural gas (NGCC); PC: Pulverized coal combustion; SC: Supercritical coal combustion; IGCC: Integrated
gasification combined cycle using coal [24].

negates the need for freshwater withdrawal, while at the same time coal-fired power generation naturally raise further concerns for
producing highly concentrated flue streams of near capture ready CO2. increased CO2 emissions.
It is generally agreed that increased use of fossil fuels since the start The increased use of natural gas as the low carbon resource for
of the industrial revolution in the mid 18th century has been largely electrical power generation is expected to continue as long as the natural
responsible for the dramatic rise in the anthropogenic CO2 concentration gas prices remain competitive with coal. The preference for natural gas
in the atmosphere. Although the percent shares of fossil fuels in global is partly due its smaller carbon footprint (~ 470 gCO2/kWh of electricity
electricity generation have been steadily falling in recent years due produced), which is nearly two times less than CO2 emissions from coal-
mainly to the rapid expansion and deployment of renewable sources, the fired power plants that generate 700 to 1000 gCO2/kWh. Around the
actual amounts of fossil fuels consumed globally for power generation as globe, air-based combustion has been the backbone of thermoelectric
well as for other energy needs have been increasing and this trend is power generation from fossil fuels, but the resulting plant level effi­
expected to continue into the mid-century, as depicted in Fig. 4. ciencies have been dismally low. For example, the average efficiencies of
Accordingly, energy-related emission of CO2 is also expected to increase US thermoelectric power generation plants tested and monitored by the
into the foreseeable future, as illustrated in Fig. 5. US Department of Energy over a 5-year period are around 33% for coal-
Even as the prospect of imminent climate change is looming high, and natural gas-fired plants [29]. As air is employed for fossil fuel
many countries are electing investments in fossil fuels over renewables combustion in these thermoelectric power plants, the stack gasses from
and green energy [4,25]. In fact between 2010 and 2018, whereas U.S. natural gas and coal fired plants contain, respectively, about 3% to 5%
and Europe decommissioned nearly 140 GW of coal-fired power ca­ CO2 and 10% to 15% CO2, with the remaining made up largely by ni­
pacity, India and China respectively added 91 GW and 342 GW of trogen. In the case of natural gas power plants, water vapor is also a
coal-fired power capacity within the same period [27] for a net gain of significant constituent in stack gasses. Due to similarities in their mo­
293 GW. In addition, China is actively building more than 140 coal-fired lecular properties, separation and capture of such lean concentrations of
power plants in countries along their Belt and Road Initiative tran­ CO2 from N2 is difficult and requires expensive and energy intensive
scending China, Eurasia, and Africa [28]. Such heavy investments in post-separation processes.

Fig. 4. Global trends in electrical power generation by source. While the share of renewables in the electricity mix is projected to increase, the amounts of natural gas
and coal consumed for power generation are also expected to increase through mid century [26].

3
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 5. Energy-related CO2 emissions by fossil fuel type [26]. (Vertical axis is in units of billion metric tons of CO2).

There is, however, an opportunity to significantly reduce the scale of diversity of these options necessarily limit the scope of the present
energy penalty and financial cost of carbon capture by phasing out air- article to review select systems, materials and technologies that offer the
based combustion, while rapidly transitioning into oxygen-based potential for practical use. It provides a structured overview without
advanced technologies such as high temperature fuel cells [22,30], advocating or recommending any one particular technology option over
some of which are discussed in the present article. They promise not only others, as this author believes in the need for a multi-prong approach to
higher efficiencies, but more importantly, possibility to facilitate pro­ address and mitigate CO2 emissions, as one size technology does not fit
duction of highly concentrated CO2 flue streams that would enable easy all. Accordingly, this review presents relevant fundamental concepts,
and cost effective capture from these point sources. At the same time, advances in materials and processes for carbon capture, discusses ad­
oxygen-based power generation can buy time to rapidly develop and vantages and shortcomings, and highlights potential risks of individual
build up requisite energy storage capacities at the TWh scale to enable technological options, while providing insights for active researchers
faster rates of penetration of renewables into the electrical grid. and decision makers. Also, tutorial materials for the uninitiated are
There is a wide range of technology options for carbon capture, presented, where appropriate.
utilization and storage (CCUS). Fig. 6 summarizes many of these ap­ Besides the 2015 National Academy of Sciences report [32], there
proaches and illustrates the possible pathways as well as interlinks. have been excellent reviews on various aspects of CCS provided in the
Significant advances in both scientific understanding and technological literature. A recent article provides an excellent and extensive review of
development of various options for CO2 capture, storage, and utilization carbon capture, utilization and storage (CCUS) technologies and their
have been achieved. Many of these topical areas, but not all, are pre­ techno-economics with focus on commercialization and integration of
sented and discussed in this article. At the same time, the breadth and CCS into the electricity system for decarbonization [33], while the

Fig. 6. A comprehensive portrayal of the multiplicity of technological options and processes for carbon capture, utilization, and storage towards achieving net-zero
or negative emissions [31].

4
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

pathways to achieve net-zero emission energy systems across a broad GHGs including CO2 and their ability to absorb energy i.e., radiative
range of energy services and industrial sectors were presented and dis­ efficiency, differ widely by several orders of magnitude, as depicted in
cussed in a recent article [34]. Another review considered a combination Table 1. Individual contributions of CO2 and other GHG gasses on
of two or more different CCS technologies and suggests that hybrid climate are generally expressed in terms of their “global warming po­
processes offer advantages in energy penalty and CO2 recovery when tential” (GWP). This is a measure of the amount of energy absorbed by 1
compared to stand alone systems [35]. Capture and reuse of CO2 for ton of CO2 over a given period of time, usually considered over a span of
supercritical cycles was reviewed with an eye for power generation and 100 years, although some databases also use 20 years. By definition, CO2
industrial applications [36]. These and others [37–40] are valuable has a GWP = 1 regardless of the time period. The lifetime of CO2 in the
contributions to our knowledge base in appreciating the complexity and atmosphere is difficult to ascertain accurately, as portion of the CO2
enormity of the challenge. While complementing the recent reviews, the emissions is captured by terrestrial as well as biological sinks.
present article provides a broad overview of processes and materials for Although not the most potent greenhouse gas for its GWP, CO2
CO2 capture from electrical power generation operations, and distin­ constitutes the largest volume GHG that greatly impacts global climate
guishes itself from most of the previous reviews by the breadth and change by absorbing outgoing infra-red radiation from the earth surface
variety of technological and materials options it covers. and re-emitting part of it back to Earth. In other words, it serves as a
The present article also promotes that placing equal emphasis on thermal shield placed around the planet. Indeed, CO2 is responsible for
capturing the ‘output’ CO2 emissions deserves as much attention as fully 65% (or, 1.7 W/m2) of the total global downward radiative forcing of
decarbonizing the ‘input’ sources for energy and electricity production. 2.6 W/m2 for all greenhouse gasses collectively [45]. Hence, carbon
This author believes this strategy can contribute greatly to a smooth capture and storage (CCS) is central to curb the amount of CO2 released
transition into renewables by allowing time for storage technologies as into the atmosphere and mitigate climate change [46].
well as capacities to catch up with the pace of penetration of renewables Once released, however, CO2 stays in the atmosphere for a long time.
into the electric grid. In that regard, the article discusses oxygen-based The lifetimes, radiative efficiencies and relative GWP values of a select
combustion and conversion technologies, and introduces high temper­ group of greenhouse gasses [43,44] are summarized in Table 1. The
ature fuel cells both for carbon abatement and capture, as well as for 20-year, 100-year, and 500-year GWP columns, represent energy
electrochemical separation of CO2 from lean flue streams. Furthermore, absorbed over 20, 100 or 500 years. For those gasses such as CH4 that
it presents the status of global CO2 emissions and updated statistics to have a shorter lifetime in the atmosphere compared to CO2, their
emphasize the scale and magnitude of the challenge, identifies many of 20-year GWP values are generally magnified, while for many of the
the knowledge and technological gaps, and provides insights for op­ fluorocarbons, chlorofluorocarbons, hydrofluorocarbons, hydro­
portunities to achieve clean electricity. chlorofluorocarbons, perfluorocarbons, and sulfur hexafluoride that
It should be noted, however, that capturing CO2 from industrial have longer lifetimes than CO2, their 20-year GWP values are signifi­
processes such as cement and steel production [41] and from distributed cantly reduced [47]. The complex interplay between CO2 and the other
sources such as transportation, as well as engineering and greenhouse gasses in the atmosphere with their respective radiative
techno-economic analyses of different technology options all lie outside forces and processes collectively result in the greenhouse effect.
the scope of this overview article, although relevant information and Interestingly, water vapor in the atmosphere accounts for nearly
published cost figures, where available and appropriate, are provided 60% of the greenhouse effect [43,44] and facilitates the largest positive
for completeness. feedback loop, i.e., increased temperatures also increase the amount of
water vapor that the atmosphere can hold. Unlike other greenhouse
2. CO2 emissions, sources and sinks gasses including CO2, however, human activity has very little impact on
the amount of water vapor uptake by the atmosphere. Also, both water
2.1. Radiative impact of atmospheric CO2 vapor and ozone had short lifetimes on the order of days and weeks in
the atmosphere. Curiously, the impact of CO2 (and water vapor) on the
CO2 is an oxidation product of carbonaceous fuels and chemicals, radiative effect of sunrays was first recognized and reported as early as
and constitutes nearly ~3/4th of the global greenhouse gasses (GHG) in 1856 [48]. There is now general agreement among the scientific
emissions [42], as shown in Fig. 7. Fossil fuel combustion makes up a community that anthropogenic CO2 emissions indeed play a major role
large portion of the global CO2 emissions. The atmospheric lifetimes of in climate change [49–51].

2.2. Global CO2 emissions

There are multiple sources of CO2 emissions, including

Table 1
Atmospheric lifetimes, radiative efficiencies, and global warming potentials
(GWP) of select greenhouse gasses over 20, 100 and 500 years. Data are
compiled from IPPC 2007 Fourth Assessment Report [43,44].
Gas Lifetime Radiative 20-year 100-year 500-year
(years) Efficiency GWP GWP GWP
(W.m-2.ppb-1)

CO2 – 1.4×10–5 1 1 1
CH4 12 3.7×10–4 72 25 7.6
N2O 114 3.03×10–3 289 298 153
CCl3F 45 0.25 6730 4750 1620
CClF3 640 0.25 10,800 14,400 16,400
CCl4 26 0.13 2700 1400 435
CHClF2 12 0.2 5160 1810 549
CHF3 270 0.19 12,000 14,800 12,200
CH2F2 4.9 0.11 2330 675 205
SF6 3200 0.52 16,300 22,800 32,600
Fig. 7. Global greenhouse gas emissions by gas type (based on IPCC 2014
CF4 50,000 0.10 5210 7390 11,200
report) [42].

5
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

anthropogenic CO2 emissions resulting from human activities, and those


from natural sources and causes such as wild fires. Human activity is
considered to be responsible for approximately 1.0 ◦ C warming of our
planet above the pre-industrial period, and limiting global temperature
rise to 1.5 ◦ C will require achieving net zero carbon by 2050 as well as
removing 100 to 1000 Gt of CO2 from the atmosphere over the 21st
century [13]. Achieving such unprecedented scale of CO2 removal is a
daunting, if not a formidable task. Emissions from natural causes are
obviously more difficult to control and manage, but man-made emis­
sions offer opportunities for capture. In that regard, emissions from
stationary point sources are the preferred targets for carbon capture than
those from distributed sources such as transportation. Regardless of the
type of CO2 emissions sources, most man-made emissions employ
air-based combustion of fossil fuels in their process chains. The resulting
product streams released to the atmosphere necessarily contain
considerable amounts of nitrogen, and also show variability in their CO2
concentrations. This is summarized in Table 2, which presents the
average concentrations and partial pressures of CO2 in the flues streams
of various industrial production and power generation plants. Typically,
Fig. 8. . Contributions of individual sectors to the global greenhouse gas
the partial pressures of CO2 in flue streams of power generation plants
emissions,(based on IPCC 2014 report) [9].
are significantly smaller than for industrial production plants, which
render CO2 separation comparatively more difficult and costly from
power generation than from industrial flue streams [52]. commensurate growth in CO2 emissions. Earlier reports from the Global
Nearly 70% of global CO2 emissions is due to fossil fuels used pri­ Carbon Project [54] and others [55,56] indicated that the global CO2
marily for power generation, transportation, and various industrial emissions have shown a stark increase in 2017 and 2018 after remaining
processes including cement and steel production. The energy sector has largely flat during the 2014–2016 period as shown in Fig. 9, which
been the largest contributor responsible for nearly 1/3rd of the global differentiates the relative contributions of advanced economies
anthropogenic greenhouse gas (GHG) emissions. This is shown in Fig. 8, compared to the rest of the world during the period 1990–2019. In this
where both distributed (e.g., transportation) and stationary point source context, advanced economies are represented by the European Union in
(e.g., electricity generation, industry) activities contribute significantly addition to 13 other countries. Fig. 9 also indicates that while energy
to total emissions. Such diversity of emission sources and technology related CO2 emissions fell by 3.2% in 2019 in advanced economies, Asia
types necessarily demand different approaches to properly address CO2 dominated by China and India, was responsible for nearly 80% of the
capture and storage. emissions growth in the rest of the world (see Fig. 1 above).
Capturing CO2 from distributed sources such as transportation, Interestingly during the COVID-19 lockdowns in 2020, the pre­
which account for nearly 1/6th of the global emissions, obviously poses liminary data had predicted that the world’s CO2 emissions would drop
greater difficulties than capturing from massive point sources. A diesel by ~7% relative to 2019 emissions [57]. Indeed during 2020, world’s
engine vehicle emits nearly 2700 kgCO2/m3 of diesel, while a gasoline energy demand dropped 4% and resulted in ~2 GtCO2 (or, 5.8%)
engine vehicle emits nearly 15% less, or 2300 kgCO2/m3 of gasoline reduction in global CO2 emissions to 31.5 GtCO2, but the downward
[53]. In other words, a typical passenger vehicle driven 18,200 km trend is expected to rapidly reverse itself to rebound by 4.8% increase in
annually with an average fuel efficiency of 9 kilometers per liter of CO2 emissions by the end of 2021 [58]. Hence, the impact of the 2020
gasoline emits 4.7 tCO2/year. With more than 281 million vehicles on pandemic on CO2 emissions has been rather limited [59].
the road in 2020 in the US and nearly 1.4 billion in the world, the Despite public pressure for environmental and climate change con­
resulting annual emissions are nearly 1.3 GtCO2 and 6.6 GtCO2, cerns, the world’s addiction to fossil fuels is not easy to shake off. The
respectively. The latter makes up close to 20% of the global CO2 emis­ Production Gap Report 2020 by the UN Environment Programme [60]
sions. Capturing such large quantities of CO2 in gigatons from such small highlights the large gap between planned fossil fuel production levels for
and distributed emitters pose significant technical and logistical chal­ countries and the global levels necessary to curb the temperature rise,
lenges far above and beyond those posed by capturing it from point and indicates that the projected growth trajectory points to increased
sources. Naturally, rapid electrification of the transportation sector production in global fossil fuels by 2030 that will be more than 120% of
greatly helps reduce these emissions. Increased global demand in energy what would have been required to limit the temperature rise to 1.5 ◦ C.
heavily dominated by fossil fuels (see Fig. 4) has resulted in Indeed, the CO2 concentration in the atmosphere, which was around 280
ppmv during the start of the industrial revolution in 1750′ s, has been
steadily rising at an alarming rate since then. Indeed, monthly average
Table 2
measurements at Mauna Loa station in Hawaii indicated that the CO2
Typical CO2 concentrations and partial pressures in the flue gas streams of
various power generation and industrial production plants [52].
level in the atmosphere has now reached 419.05 ppmv in April 2021,
which represents a 2.6 ppmv increase from 416.45 ppmv measured in
Plant or Industry Type Ave. CO2 conc. PCO2
April 2020 [61]. This is depicted in Fig. 10. What is more disturbing is
(v%) (MPa)
the fact that the annual mean growth rate in atmospheric CO2 concen­
Coal-fired power generation 12 to 14 0.012 to 0.014 tration has been steadily increasing from 0.9 ppmv/yr in the 1960′ s to
Natural gas turbine power generation 3 to 4 0.003 to 0.004
Oil-fired power generation 11 to 13 0.011 to 0.013
1.5 ppmv/yr in the 1990′ s to nearly 2.5 ppmv/yr in 2010′ s [62]. The 5th
Integrated gasification combined cycle 12 to 14 0.012 to 0.014 Assessment Report of the Intergovernmental Panel on Climate Change
Cement production 14 to 33 0.014 to 0.033 (IPCC-AR5) had concluded that rapid decarbonization and trans­
Steel production (blast furnace) 20 0.040 to 0.060 formation of the electricity generation sector into low-GHG alternatives
Ammonia production 18 0.5
is required to stabilize the atmospheric levels at the 430 to 530 ppm
Ethylene oxide production 8 0.2
Hydrogen production (steam reforming of methane) 15 to 20 0.3 to 0.5 CO2e (i.e., CO2 equivalent) levels by 2100 [63]. Note that CO2e repre­
Methanol production 10 0.27 sents the same amount of time-integrated radiative forcing over a given
period of time by greenhouse gasses compared to CO2 emissions. Earlier,

6
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 9. Energy related global CO2 emissions (1990-2019), which flattened in 2019 around 33 GtCO2 following 2 years of increase between 2016-2018. Lower bars are
for advanced economies, and upper bars indicate the rest of the world [6].

Fig. 10. Recent readings of CO2 concentration in the atmosphere, measured at the Mauna Loa Observatory in Hawaii [61]. Note the gradual acceleration in the rate
of increase in the last several decades.

the 2007 IPCC report had recommended that atmospheric CO2 levels lacking.
should be stabilized around 450 ppmv by mid-century in order to avoid Left unchecked, CO2 emissions pose unprecedented risks for the
series risks for the health and well-being of our planet’s ecosystems [43, delicate ecosphere of our fragile planet. Consequential impacts of the
44]. By either count, the atmospheric CO2 concentration is fast rapid build-up of atmospheric CO2 are generally associated with acidi­
approaching the critical zone identified by these IPCC reports not to fication of the oceans, warming of the planet, disruption of ecosystems,
transcend in order to stabilize climate change. and climate change, including extremes of weather events and climatic
The recent United Nations Emissions Gap Report of 2019 [5] sent a conditions such as extended droughts, severe storms and floods, rapid
similar warning and cautioned the world is on course for a 3.2 ◦ C tem­ melting of continental ice [65,66], and sea level rise, all of which may
perature rise with nearly 55.3 GtCO2e of emissions in 2018, of which alter or harm the earth’s ecosystems in possibly irreversible ways.
37.5 GtCO2 was from fossil fuel emissions. The 2019 report also sends a Worse, we may be fast approaching a dangerous threshold where
dire warning to the global community to take immediate action in reversing the emissions trend simply by capturing CO2 from point and
dramatically reducing total emissions by 25% and 55% lower (or, 15 distributed emission sources may not be sufficient to eliminate the
GtCO2e and 32 GtCO2e) than 2018 values in order to limit global possibility of irreversible change [67]. Unless drastic measures were
warming, respectively, to below 2 ◦ C and 1.5 ◦ C by the year 2030. It implemented urgently to reverse the rising trend in CO2 emissions, ef­
points out that there is still opportunity to limit the global temperature forts to adapt to new environmental and climatic constraints are ex­
rise to 1.5 ◦ C, a target also recommended by the Intergovernmental pected to face difficult choices and serious challenges.
Panel on Climate Change (IPCC) 2015 and 2018 reports [10,13,64] If nations do not act but wait until 2025 to take action, more drastic
provided that all countries collectively boost their current commitments measures including 15.5% reduction in emissions per year will be
by 5-fold and commit to cutting their emissions by 7.6% every single required. Such inaction will greatly jeopardize if not make it very
year between 2020 and 2030 in order to achieve the 25 GtCO2e emis­ difficult and costly to achieve the 1.5 ◦ C target. The UN Emissions Gap
sions limit. Needless to say, this imminent and urgent threat is yet to be Report 2019 also indicates that limiting global temperature rise to
acted upon by the global community, and commitments are sadly within 2 ◦ C by 2100 requires the greenhouse gas (GHG) emissions not to

7
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

exceed 41 GtCO2e by 2030, while 1.8 ◦ C rise and 1.5 ◦ C will require oxygen-based advanced power generation technologies [22], as well as
larger reductions keeping total emissions below 34 GtCO2e, and 25 better management of terrestrial sources and sinks for CO2 [72,73].
GtCO2e, respectively. The report also estimates that policies commen­
surate with achieving 1.5 ◦ C rise will demand a significant annual in­
2.3. Global CO2 sources and sinks
vestment between $1.8 trillion and $3.8 trillion per year by the global
community over the 2020–2050 period to upscale their energy systems.
There are both man-made (i.e., anthropogenic) and natural sources
Unfortunately, the pace and the scale of measures taken by the global
for CO2 emissions, and human activity contributes heavily into that
community have not been sufficient to meet these targets [68]. 2020
outcome. Any strategy to tackle the emissions challenge should also
Cancun pledges by the G20 countries to collectively reduce their emis­
involve more effective and efficient management of terrestrial sources
sion by 1 GtCO2e per year is severely lagging behind with many
and sinks for CO2. Less than half of the total CO2 emissions remain in the
including the U.S.A., Canada and the Republic of Korea missing their
atmosphere, while the rest is taken up and removed by oceans and the
targets. In a similar manner, nationally determined contribution (NDC)
terrestrial biosphere. Besides technological routes to capture CO2, nat­
targets for emission reductions from major economies have largely
ural and biological pathways for carbon fixation and storage including
lagged behind as indicated in Fig. 11, which compares the relative
forestation and photosynthesis, ocean fertilization and hydration (i.e.,
progress of various countries towards meeting their reduction
hydrate formation), and mineral carbonization should also be exploited.
commitments.
However, discussion of these subtopical areas falls out of the scope of
Developing not only technologies, but also policies to reduce CO2
this article, and is not covered here.
emissions are complicated further by indirect factors and externalities.
During the period 2000–2018, CO2 emissions into the atmosphere
Emissions in one geographical location do not stay in that region but
from the burning of fossil fuels rose by 24.5 Gt of CO2 (or, 6.7 Gt of C).
spread across the globe. Moreover, emissions are also intimately incor­
Indeed, the 2019 United Nations Emissions Gap Report [74] indicated
porated into the vast scale of international trade. As CO2 production in
that energy related CO2 emissions primarily due to fossil fuels use have
one country or location provide the goods and services, as well as in­
reached 37.5 Gt of CO2 (or, 10.2 Gt of C) by the end of 2018, and made
dustrial or consumer products in another country, this supply chain has
up more than 2/3rd of the total global GHG emissions of 55.3 GtCO2e.
a significant impact in accurate accounting for regional CO2 emissions
Wildfires constitute another major source for CO2 emissions, and
[69,70].
contributed an additional 7.3 to 9.5 Gt of CO2 (or, 2.0 to 2.6 Gt of C) into
It is clear that rapid mobilization of a concerted global effort is
the atmosphere [75].
required to achieve the objectives outlined in the 2018 IPCC-SR15
As depicted in Fig. 12, there are natural sinks that offset the CO2
preliminary report [10] as well as in the 2015 Paris Agreement. The
emissions from sources. As CO2 has appreciable solubility in water,
recent meeting of the Group of Seven (G7) countries on June 11–13,
oceans help capture significant amounts of CO2 by absorption, ranging
2021 in Cornwall, UK, provided a glimmer of hope for the globe in this
between 5.1 and 15.0 Gt of CO2 annually (or, 1.4 to 4.1 PgC/yr) [75].
regard, where the G7 leaders made commitments to cut their collective
Increased emissions force higher concentrations of dissolved CO2 in the
emissions by 50% by 2030 and achieve net-zero emissions before 2050
oceans to maintain equilibrium with the increasing concentrations in the
in order to limit the rise in temperature to 1.5 ◦ C [71]. To achieve these
atmosphere. Higher concentrations of dissolved CO2 decrease the pH
goals at the global scale, multi-throng strategies must be implemented
and alter the biological ecosystem of the oceans. A notable impact of
and deployed rapidly in order to reduce atmospheric carbon levels.
lower ocean pH is the destruction of the corrals, whose livelihood de­
These will most likely involve several key ingredients including carbon
pends heavily on the carbon balance.
capture and storage, increased energy efficiency, rapid transformation
Land is another major sink where CO2 is carbonized or mineralized to
of energy supply from fossil fuels to renewables, rapid transitioning of
carbonates, and adsorbed by soil. Also, the earth’s biosphere removes
fossil fuel-based thermoelectricity production from air-based to
significant quantities of CO2 from the atmosphere via photosynthesis.

Fig. 11. Required greenhouse gas reductions for the G20 countries for their nationally determined contributions (NDC) indicate that most countries lag behind and
urgently need to implement additional cuts in emissions by 2030 to achieve their NDC targets under unconditional (left) and conditional (right) scenarios [11].

8
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

3. Capture and removal of CO2

CO2 capture presents multitude of challenges, as there are widely


differing emission sources with different characteristics, compositions,
volumes, spatial and temporal distributions, as well as different tech­
nology options and materials for CO2 separation from flue streams [37,
40,77–80].Since the industrial revolution, it is estimated that more than
2040±310 GtCO2 has cumulatively been added to the atmosphere be­
tween the years 1750 and 2011 [81]. To accomplish emission reductions
at the gigaton scale requires commensurate industrial capacity, which
we do not have yet. The case in point is the fact that, by one estimate by
the American Physical Society [45], it is necessary to capture and
remove nearly 7.8 GtCO2 from the atmosphere in order to reduce at­
mospheric concentration of CO2 by only 1 ppmv. On the other hand, the
2015 National Research Council report indicates that 1 ppmv reduction
of CO2 concentration in the atmosphere requires removal of 18
GtCO2/yr [32]. The disagreement in the required removal amounts may
in part be due to differences in the assessment methodologies and
related assumptions used for the CO2 uptake and release fluxes from
different sources and sinks.
The average annual upward growth of atmospheric CO2 concentra­
Fig. 12. The annual global CO2 budgets between January 2000 and December
tions in the last 60 years is shown in Fig. 14, where a net increase from
2018, depicting natural and man-made sources and sinks for carbon. During
an average of 2.0 to 2.5 ppmv in last decade is related to CO2 emissions
this period, CO2 emissions from fossil fuel combustion rose from 6.7 PgC/yr to
10.2 PgC/yr (1 Pg of C = 1 GtC = 3.67 PgCO2). Four types of surface-to- of 35.1 and 37.5 GtCO2 [83]. Although the remnant CO2 concentration
atmosphere exchange processes are color coded in the figure for easy identifi­ in the atmosphere is a result of complex and dynamic processes and
cation. Tan indicates fossil fuel emissions, green represents terrestrial biosphere exchanges among multitudes of CO2 sources and sinks, a simplistic
flux excluding fires, red denotes direct emissions from fires, and blue indicates approximation suggests that bringing down the atmospheric CO2 con­
gas exchange between air and sea. Negative emissions denote CO2 removal centration even 1 ppmv will likely require annual removal of multiple
from the atmosphere, while the thick black lines indicate the net surface ex­ gigatons of CO2 from air. This presents an immense industrial challenge.
change computed from the sum of the four fluxes [75]. As an equivalency benchmark to emphasize the magnitude of the
problem, offsetting, i.e., avoiding, the release of 1 GtCO2 into the at­
Hence, better management of biomass and land use such as reforestation mosphere would correspond to installing 750 GW of solar PV, 270,000
and afforestation, crop selection and productivity are important con­ of 1 MW wind turbines, 136 of 1 GW nuclear power, 273 of 500 MW
siderations [76]. Although there are wide variations in the spatial and zero-emission coal-fire plants, 1000 of Norway’s Sleipner offshore CO2
temporal distribution of land sinks, terrestrial biosphere also helps storage sites, and reforestation of 900,000 km2 of barren area almost the
remove 9.1 to 17.6 Gt of CO2 (or, 2.5–4.8 Gt C) annually from the at­ combined sizes of Germany and France [84]. Aside from the technical
mosphere. Collectively, these natural terrestrial sinks remove nearly and economic difficulties of CO2 capture, the scale of the magnitude of
50% of the anthropogenic CO2 from the atmosphere [75]. A more this challenge presents an unprecedented industrial built-up and finan­
detailed budget for the carbon cycle of man-made and natural CO2 cial undertaking. Needless to say, reducing the atmospheric concentra­
sources as well as sinks is presented in Fig. 13, which indicates an tions from today’s 419 ppmv to the pre-industrial level of ~280 ppmv
average accumulation of 4 GtC/yr (or, 14.67 GtCO2/yr) in the atmo­ (or, by 134 ppmv) by active removal of CO2 from the atmosphere is an
sphere. It should be noted that another recent study on the global carbon impossibly formidable task.
budget indicates that the annual average accumulation of CO2 in the The CO2 emissions at the tens of gigatons scale underline the stark
atmosphere has been 5.1 GtC/yr (or, 18.70 GtCO2/yr) during 2010 and reality that any technology for carbon capture and storage (CCS) has to
2019 with a small uncertainty of only ±0.02 GtC/yr [57]. The be deployed at comparable scale in order to make a dent in the net
discrepancy between the two carbon budget studies highlights the dif­ emissions released to the atmosphere [72,73]. The magnitude of the
ficulty and complexity of accurate accounting for emissions. challenge to build the industrial capacity that can handle the volumes
Interestingly, gross CO2 emissions from natural processes occurring necessary for capturing CO2 is daunting and is often overlooked. Table 3
on lands and oceans are nearly 20 times the volume of anthropogenic compares the scales of magnitude of 2018 energy related CO2 emissions
CO2 emissions from fossil fuels and other human activities [32]. This from fossil fuel sources with the world’s total existing production ca­
requires that the efforts for better management of lands and oceans pacities of the largest volume chemicals and plastics, and provides a
should go hand-in-hand with efforts to curb anthropogenic CO2 emis­ sobering snap shot of the severity of the challenge the world faces. The
sions. As expected, the individual contributions and roles of natural wide chasm between our largest industrial capacities currently in place
terrestrial processes and man-made emissions in the carbon cycle and what would be needed for capturing the vast volumes of CO2
display wide variations in both their geographic and temporal distri­ emissions in the multi-gigaton scale clearly highlights the enormity of
butions. Nevertheless, Fig. 13 illustrates the complex and intricate dy­ the problem. Similarly, the total global volume of petrochemicals in
namic balances and fluxes among the global carbon sources and sinks. 2021 is expected around 310 Mt, not significantly different from com­
There is concern, however, that this dynamic equilibrium may not be mon chemicals in Table 3. Outside the chemical industry, global pro­
robust enough to tolerate sudden and massive disruptions or in­ duction capacities of only liquid fuels from petroleum (~6 Gt in 2019),
terventions to the carbon cycle. Indeed, a recent report by the National cement (4.1 Gt in 2018) and steel (1.8 Gt in 2018) industries are
Research Council (NRC) cautiously points to the possible unintended commensurate with the multi-gigaton capacities needed for CO2
consequences of massive carbon capture, where significant quantities of capture.
CO2 removal from the atmosphere may for example lead to CO2 out­ Given also the widely varying nature and sources for CO2 emissions,
gassing from the oceans [32]. complementary multi-prong strategies are required to achieve net
removal of CO2 from the atmosphere at the gigaton scale. In this regard,
it was recently suggested that a systems approach, including effective

9
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 13. Global carbon cycle indicating annual fluxes among major sources and sinks in PgC, or Gt of carbon (1 PgC = 3.67 Gt CO2), where positive numbers indicate
accumulation. Arrows indicate fluxes in PgC/yr, while boxes denote stocks or reservoirs in PgC. Black numbers and arrows indicate exchange fluxes and reservoir
masses estimated prior to 1750, and red numbers on arrows denote annual fluxes of anthropogenic carbon over the industrial period between 1750 and 2011. Red
numbers in reservoirs denote the cumulative accumulation of anthropogenic CO2 over the industrial period [32].

Table 3
Comparison highlighting the large deficiency in the world’s largest installed
chemical production capacities for industrial chemicals against the energy-
related global CO2 emissions (** from [85]).
2018 Global Production
(million tons)

Global energy-related CO2 emissions [**] 35,319


Coal [**] 15,041
Liquid fuels [**] 12,980
Natural gas [**] 7,298
Chemicals Production
Sulfuric acid 266
Ammonia 168
Ethylene 180
Propylene 120
Plastics Production (total) 359

Fig. 14. Annual average increase in atmospheric CO2 concentrations measured this magnitude and diversity, a concerted effort to develop a range of
at the Mauna Loa observatory in Hawaii, U.S.A. [82]. advanced technologies is required both to capture CO2 as well as to
avoid its production. The latter can be achieved by improved efficiencies
management of CO2 exchange among the atmosphere and lands and in the use and oxygen-based conversion of fossil fuels for energy and
oceans, exploiting the natural biological carbon cycle including refor­ electricity production [22]. Additional pathways must include devel­
estation and crops, large scale carbon capture and storage (CCS), and oping advanced technologies for net-zero emissions, carbon-neutral
reutilization of CO2 and conversion into useful chemicals and fuel fuels and energy sources, and rapid transformation of the energy sup­
products may collectively help ease the burden on requisite industrial ply from fossil fuels to predominantly renewable energy sources.
capacities for CO2 capture [31,73]. To respond to a global challenge of At the same time, development of large scale carbon capture and

10
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

storage (CCS) options and technologies have been struggling to gain plant efficiency when CCS is implemented. Similar findings were re­
public acceptance and investments towards mitigating climate change. ported by another DOE study, which compared plant efficiencies and
This is partly due to divergence of public opinions regarding the feasi­ cost of electricity for various advanced coal and natural gas-based
bility, cost, reliability, and long-term risks of CCS. Current lack of thermoelectric power generation with CCS [88].
accepted and widely implemented mitigation policies and a global Primarily, CCS consists of three major steps, namely, CO2 capture,
pricing structure for carbon emissions further complicate the picture and transport (in a compressed state), and long-term storage. Among the
in fact, partially and covertly subsidize the low cost of fossil fuel-derived three, capturing of CO2 is by far the most costly step in the overall value
energy with potentially drastic consequences for the planet. chain, and depending on the emission source and other factors, it may
On the financial side, the World Energy Outlook 2016 report [86] account for 50% to 90% of the overall cost of CCS [89]. This assessment
suggested that limiting global temperature rise to within 2 ◦ C above the is supported by other studies that estimated the cost of the CO2 capture
pre-industrial period with 50% chance of success would require in­ step might exceed 70% to 80% of the total cost of CCS [90,91]. One of
vestment of at least $44 trillion. About 60% of the $44 trillion was ex­ the most advanced technology options for industrial scale CO2 capture
pected to go to extraction, production and supply of fossil fuels, as well involves chemical separation using liquid amine-based solutions. This
as the use of oil, gas, and coal for power production, while only 20% for also serves as a benchmark for comparing other technology options.
renewable energy sources. An additional $23 trillion would be needed However, the amine-based process is costly and energy intensive. It also
for efficiency improvements in energy systems, services and products. has other disadvantages including the chemical stability and environ­
Even with such an enormous investment, the report expected that mental impact of the amine solution during thermal cycling as well as
hundreds of millions of people around the world would still not have the corrosion of the vessel over long use.
access to basic energy services for their needs. Another important consideration in CO2 capture is the concentration
In all its forms, carbon capture and storage (CCS) is a costly and (or partial pressure) of CO2 in the flue stream. As expected, the cost and
energy intensive process. As expected, a 2010 U.S. Department of En­ energy penalty for capture increases with decreasing CO2 concentration
ergy (DOE) report from the National Energy Technology Laboratory in the product stream (see Fig. 18 below). Additionally, the temperature
(NETL) has concluded that CCS adversely impacts the overall system and chemical make up of other constituents in the flue stream also
efficiency, plant capital cost and levelized cost of electricity (LCOE) of impact the suitability and efficacy of the capture technology. These
much of the fossil fuel-based advanced thermoelectric power production considerations may be particularly important for capturing CO2 from
in a substantial way [87]. Key findings of this study are presented in massive point source flue streams such as fossil fuel-based power plants,
Fig. 15, which indicates significant increases on both the plant capital cement production, and iron and steel plants, where the product gasses
cost and the levelized cost of electricity, while substantially decreasing typically exit the process stream at significantly elevated temperatures,

Fig. 15. Adverse impact of carbon capture and storage (CCS) on plant efficiency, capital investment, and levelized cost of electricity for thermoelectric power
generation by integrated gasification combined cycle (IGCC) and integrated gasification fuel cell (IGFC) technologies employing advanced processing equipment such
as pumps, turbines and membranes. The figure clearly demonstrates the substantial advantage of incorporating fuel cells in power generation and their beneficial
impact on both cost and performance [87]. Filled squares – IGCC; Open squares – IGFC; Blue symbols – with CCS; Red symbols – without CCS.

11
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

which would require cooling. Capturing especially from flue streams There are several technology options for CO2 capture from point
with low CO2 concentrations presents major technical challenges to sources as well as directly from air, as depicted in Fig. 16. Needless to
achieve high yield at high efficiency. Furthermore, the capital costs and say, climate change mitigation measures must also involve CO2 capture
energy penalties for CO2 capture show wide variations depending on the from other point sources including industrial processes such as cement
type of power plant and fuel, geography, volume and the CO2 content of and steel manufacturing, petrochemical plants, as well as distributed
the flue stream, and the type of capture technology adopted. The cost sources such as transportation in all of its forms. Generally accepted
estimates are also impacted by the end use. For example, it was esti­ carbon capture categories are, namely, pre-combustion, post-combus­
mated that the total cost of CCS from power plants ranges from $4.5/ tion, oxyfuel combustion, and direct air capture. The latter is self
tCO2 for coal-based integrated gasification combined cycle (IGCC) plant explanatory, as it involves capturing CO2 directly from air regardless of
with the end use of CO2 intended for enhanced oil recovery, to $72.4/ the location of CO2 emissions, while oxyfuel combustion involves the
tCO2 for a natural gas combined cycle plant with long-term storage of use of oxygen – instead of air – for burning the fossil fuel and hence,
CO2 in saline formations [92]. Similar results are reported for other yields a product flue stream that is highly concentrated in CO2, which
industrial processes including iron and steel, cement, refinery, and pulp practically eliminates the need for post separation operations. Pre-
and paper, where the costs of CCS are estimated to vary between combustion capture involves gasification of the fossil fuel into syngas
$30/tCO2 and $70/tCO2 depending on the type of emission source and (i.e., mixture of CO and H2) by reacting it with steam and oxygen, fol­
CO2 concentration [41,93]. lowed by the combustion of the syngas (or, hydrogen after a water-gas
Needless to say, CO2 capture offers significant environmental bene­ shift reaction step) in a gas turbine to generate electricity. Here, CO2
fits and helps reduce its global warming impact. A recent life cycle is separated and captured from the shifted syngas usually before the
analysis reported that the GWP for a pulverized coal (PC) power plant resulting hydrogen is combusted in a gas turbine. By contrast, post-
without CCS is 876 kgCO2e/MWh while it is 203 kgCO2e/MWh for post- combustion capture refers to collection and removal of the product
combustion capture using amine solution, and only 154 kgCO2e/MWh CO2 from the process flue stream after combustion has taken place. The
for oxy-combustion [94]. Similarly, this study estimated the GWP of present article also introduces the promising yet nascent technology of
IGCC plant without CCS to be 1009 kgCO2e/MWh, while it is 190 high temperature fuel cells that excludes nitrogen from entering the
kgCO2e/MWh with pre-combustion CO2 capture. It also concluded that process stream and allows easy capture of the product CO2.
the GWP of power plants could be reduced significantly with CCS by Given the magnitude of the global emissions challenge, a task force
nearly 63–82%, where the biggest environmental benefit is offered by report on CO2 utilization and negative emissions technologies by the U.
implantation of oxy-combustion in PC and IGCC power generation. S. -DOE Energy Advisory Board (SEAB) [31] has considered various
There are more than 300 CO2 capture and storage projects at various scenarios that offer the potential for the requisite annual capacity of ≥ 1
stages in more than 30 countries across the globe. The U.S. Department GtCO2, and estimated that investments of the order of $100 billion will
of Energy’s (DOE) National Energy Technology Laboratory (NETL) be required to achieve 1 GtCO2/yr. For such vast investments to mate­
maintains an interactive database that keeps track of these projects [95]. rialize, carbon capture must be cost-effective and needs to create value
Among these, the database indicates 93 active, 77 completed, 36 on that is bigger than its cost. Indeed, the ‘Carbon Capture Program” of the
hold, and 58 terminated projects. Other institutions offer similar data­ U.S. -DOE indicates that the cost of capture came down from $100/tCO2
bases for the status of CCS projects as well [96–98]. Unfortunately, a in 2005 with ~30% energy penalty to ~$40/tCO2 in 2020 with ~10%
large fraction of these projects have annual capacities of less than 1 energy penalty for the produced power [102].
MtCO2, which fall far short of making a dent on the multi-gigaton scale It should be noted that most projections for the future of energy and
of CO2 emissions (see Table 3). The Global CCS Institute also keeps track CO2 emissions are unfortunately based on the implicit assumption and
of some of the large-scale CO2 capture projects including oxy-fuel de facto acceptance of air-based combustion to largely prevail as un­
combustion, pre-combustion and post-combustion capture [99], and challenged status quo into the foreseeable future for power generation as
its October 2019 announcement included 51 CCS facilities across the well as for other industrial processes. However, there is no fundamental
globe, with 19 in operation, 4 under construction, and 28 in various reason or basis for this to be the case. Comparative costs may be the only
stages of development, for a combined capture capacity of only 96 argument against it. However, do the current costs provide good enough
MtCO2 annually [100]. Clearly, these are insufficiently low capacities reasons to dismiss that option? In fact, oxygen-based combustion and
and the world urgently needs much more ambitious initiatives, policies, power generation processes offer attractive options [22], which are
and projects in the gigaton scale in order to curb the rise in CO2 largely ignored by policy and decision makers. As a result,
emissions. post-combustion capture has acquired most of the attention and interest
amongst other options. Naturally, nitrogen in the air, as an inert
3.1. Strategic choices and technology options bystander throughout the combustion-to-power conversion process
chain, makes it difficult to separate out and capture the CO2. Several
Nearly half of the anthropogenic CO2 emissions are from distributed technology options for post-combustion capture are available for sepa­
sources including transportation and residential and commercial heat­ rating CO2 from flue gas streams, including physical and chemical ab­
ing, while the remaining is largely from massive point sources such as sorption in liquid solvents, adsorption on solids, mineralization (or
power plants, cement and steel manufacturing, refineries, combustion carbonation), and membrane or cryogenic separation. Criteria for
and fermentation processes. Such diversity and sheer volume of CO2 selecting the optimal process to capture CO2 requires careful consider­
emissions necessarily require investments in a wide range of technolo­ ation of the type, volume (or rate), and the CO2 concentration of the
gies to properly address CO2 capture from these divergent sources. emission source, as well as the economic and energy costs for the sep­
Clearly, one size does not fit all. Accordingly, multi-prong strategies are aration process. Among the criteria, the composition of the flue gas
needed to achieve dramatic reduction in atmospheric CO2, and these deserves critical consideration, as the CO2 content of flue gasses differ
should involve several key pathways including development of widely among stationary power generation sources. For example, the
advanced technologies for negative as well as net-zero emissions, cap­ nominal range of CO2 content in flue gasses is 3% to 5% for natural gas
ture and storage of CO2, improving energy efficiency, developing turbines but 10% to 15% for pulverized coal fired plants, while it is 40%
carbon-neutral fuels and energy sources, and rapidly transforming the to 60% after water gas shifting for integrated gasification combined
energy supply from fossil fuels to renewable sources with access to cycle (IGCC) plants.
adequate storage capacities [101]. The strategy should also include
exploiting the biological carbon cycle and effective management of
terrestrial CO2 sources and sinks including lands and oceans [73].

12
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 16. Schematic depiction of major technology options for CO2 capture from air, as well as from point source emitters such as power generation plants. The
nascent technology of power generation in high temperature carbon fuel cells is depicted with dashed lines.

3.2. Fundamental considerations theoretical minimum work, Wmin, to separate CO2 from a gas mixture in
the exhaust stream (or, air) is equal in magnitude but opposite in sign to
Separation of CO2 from air, combustion exhausts, or other process the free energy of mixing, and is given by the difference between the free
gas mixtures emitted either from stationary point or distributed sources energies of the streams (B) and (C), and the feed stream (A). For an
needs to overcome the thermodynamic barrier of demixing. A simplified isothermal and isobaric separation process, this minimum theoretical
flow diagram that summarizes the essential process steps is schemati­ work requirement can be expressed by Eq. (1).
cally illustrated in Fig. 17, where the exhaust stream (A) from an { ∑ ∑
emission source, or simply air, is fed into a CO2 capture process, which Wtheo = − RT NB XB,i ln XB,i + NC XC,i ln XC,i
produces a CO2-depleted stream (B) and a highly enriched or pure CO2 i
}
i

stream (C). Assuming ideal gas behavior for all process streams, the ∑
− NA XA,i ln XA,i (1)
i

Here, N represents the individual molar flux for streams (A), (B), and
(C), and Xi denotes the molar fractions of species i in the three streams,
respectively. Based on this 2nd Law analysis, the theoretical minimum
work requirement to recover CO2 from air (> 400 ppmv) with 99%
purity, while exhausting a depleted stream with only 200 ppm is esti­
mated to be 20 kJ/mol of CO2 [103]. This value is in good agreement
with 22 kJ/mol of CO2 reported for the free energy of mixing of CO2 in
air at 300 K [104]. The minimum theoretical work requirement de­
creases almost exponentially as the CO2 content of the feed gas supplied
to the separation unit increases. This is shown in Fig. 18, where mini­
mum requirements are presented for CO2 capture from several industrial
power generation processes. It is clear that it is energetically much less
expensive to separate out CO2 from point sources of power generation
Fig. 17. Simplified process block flow diagram for post-combustion CO2 cap­
than removing it from air. For example, the minimum theoretical work
ture either from exhaust gas mixtures of emission sources, or CO2 capture needed to capture 90% of CO2 from pulverized coal combustion (PCC)
directly from air. plant that emits a flue stream containing 12% CO2 is estimated to be

13
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

nitrogen enters the combustion process stream negating the need for
post-separation. Integrated gasification combined cycle (IGCC), inte­
grated gasification fuel cell (IGFC), oxy-combustion and chemical
looping combustion (CLC) technologies are discussed under this cate­
gory, which also includes the nascent technology of carbon fuel cells
(CFC) for coal conversion that offer efficient power generation with
nearly capture-ready CO2 production. However, as the focus of the
article is on carbon capture and not on fossil fuel-based power genera­
tion, these technologies are described in brevity for context and
completeness only, and not intended for a comprehensive review.

4.1. Air separation technologies

Separating oxygen from air is the requisite first step towards


achieving pre-combustion carbon capture from oxy-combustion and
gasification technologies [111]. The theoretical minimum energy
required to separate oxygen from air is 176 MJ/tO2 (or, 49 Wh/kgO2). In
practice, of course, the actual energy penalty may be several times
higher than this theoretical minimum [112]. For example, air separation
units (ASU) using cryogenic distillation may demand 150 to
250 Wh/kgO2, while energy demand for pressure swing adsorption
(PSA) and ion transport membrane (ITM) separation may require 250 to
300 Wh/kgO2 and 350 Wh/kgO2, respectively. Such high penalty may
consume up to 8% to 12% of the energy produced at the power plant
[113]. However, use of pure or highly enriched oxygen offers the
advantage of easy capture for fossil fuel-based thermoelectric power
Fig. 18. Minimum theoretical work requirements for capturing CO2 directly
generation as the stack flue stream is highly enriched in the product CO2.
from air or from flue streams of various power generation processes, where the
numerical ranges indicate the different CO2 purities and the percentage of CO2
Air is the most abundant and cheapest source for oxygen. Global
recovered [105]. Reproduced with permission from the Annual Review of market for oxygen production made up more than one quarter of the
Chemical and Biomolecular Engineering. world’s gas industry that generated over $87 billion in revenue in 2019
[114]. Several technology options are available to produce oxygen from
air, namely, cryogenic distillation, adsorption, and membrane separa­
158.1 kJ/kgCO2 while direct capture from air is nearly 3 times more tion [115,116], all of which require energy expenditures for separation.
energy intensive at 482.5 kJ/kgCO2 [83]. As expected, the correspond­ By contrast, high temperature (>650 ◦ C) fuel cells such as solid oxide
ing minimum work requirements are lower for 50% CO2 recovery with (SOFC) and molten carbonate fuel cells (MCFC) electrochemically pro­
134.9 and 457.8 kJ/kgCO2 for PCC and direct air capture (DAC) pro­ vide oxygen in a downhill process while generating electricity not by
cesses, respectively. Naturally, these theoretical requirements represent combustion of fuels but by electro-oxidation. This is an important
a grossly underestimated benchmark for the actual energy penalty for distinction and offers higher conversion efficiencies in fuel cells. The
separation and capture. In real life, irreversibilities and various losses
along the process pathway increase this minimum work requirement Table 4
substantially. Moreover, this simplistic 2nd Law analysis considers only Key features of air separation processes, operating principles and materials for
work, whereas actual separation processes to remove CO2 typically oxygen production. Enriched oxygen refers to purities in the range 35% to 90%
involve both heat and work. Nevertheless, it provides a useful guiding depending upon the separation process and the number of recirculation cycles
[22].
tool for CO2 capture from point and distributed sources, as well as from
air. Technology Separation Driving Force Oxygen Maturity
Method Purity Status
It was estimated that the energy penalty for CO2 separation could
amount to 25% to 40% of the energy production of a power plant and up Cryogenic Air Multi-stage Difference in 95% to Commercial
to 70% of the additional costs for CCS [106]. This study also estimated Separation distillation boiling points 97%
O2
that the energy cost for capturing and storing 80% of the CO2 emitted Adsorption on Cycling Surface binding Enriched Commercial
from the U.S. coal-fired power fleet requires 69 to 92 GW of carbon-free Solid pressure or energy difference O2
power that results in 15% to 20% reduction in the overall electricity Sorbents temperature
consumption. Increasing the power conversion efficiency to over 50% Polymeric Bulk sorption Pressure Enriched Commercial
Membranes and diffusion differential O2
can offset significant reductions in the energy penalty of CCS. As dis­
Porous Size sieving, Pressure Enriched Limited
cussed later in Section 4.6, it is indeed possible to achieve conversion Inorganic pore differential O2 availability
efficiencies of 50% or higher with the added benefit of producing membranes adsorption-
capture-ready CO2, using high temperature fuel cells for electrochemical diffusion
conversion of fossil fuels to electricity [30,107–110]. Dense Mixed ionic Pressure 100% Limited
(Nonporous) and differential O2 availability
Ceramic electronic
4. Pre-combustion carbon capture technologies Membranes transport
Selective Electrochemical 100% Not
Typically, pre-combustion capture involves an apriori step for the Ionic pumping O2 commercial
Transport
separation of oxygen from air, followed by the use of pure or highly
Carbon Fuel Selective O2 chemical 100% Not
enriched (>90%) oxygen for the fuel combustion or gasification process. Cells Ionic potential O2 commercial
This upstream separation process results in product flue gas that is made Transport gradient
primarily of capture-ready CO2 and water vapor as practically no

14
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

different technological options for air separation are summarized in


Table 4, which presents operating principles, nominal oxygen purities,
and state of commercialization of various oxygen production method­
ologies. Each offers technological and economic advantages and disad­
vantages, but not all of them are commercially available, and not all are
at the same level of maturity.

4.1.1. Cryogenic air separation


This is the most mature and reliable technology for high volume
oxygen production, and has been commercially available for many de­
cades. It is based fundamentally on the principle of Joule-Thomson
refrigeration of air that is followed up by selective multi-stage distilla­
tion via 100′ s of theoretical plates to obtain high purity (>95%) oxygen.
To achieve this, ambient air is compressed to ~1.3 MPa pressure and
expanded in subsequent steps to cool it down to the melting point of air
(− 194.35 ◦ C). Subsequent cycles allow the liquid to increasingly enrich
in oxygen (BP = − 182.93 ◦ C), while the gas phase is enriched in nitrogen
(BP = − 195.8 ◦ C). The oxygen is extracted in liquid form from the
distillation column, and is vaporized to gaseous oxygen before it is used
as feed into the intended process stream.
Cryogenic air separation requires about 240 Wh/kgO2 of energy to
produce 95% pure oxygen [117], which presents a significant energy
penalty that reduces the plant power by nearly 15%. However, utilizing
the waste heat from the ASU significantly improves efficiency by up to
20% [116,118]. Arguably, cryogenic separation is the most convenient
and mature technology for the production of large volumes of oxygen.
Typical cryogenic plant sizes may vary between 30,000 and 50,000 Nm3
of O2 produced per hour, or 9000 to 15,000 tO2/day.

4.1.2. Adsorption-based air separation


Adsorption-based air separation [115] exploits the relative heats of Fig. 19. Operating principles of (top) polymeric and microporous ceramic
membranes, and (bottom) ion transport ceramic membranes (ITM). P denotes
adsorption and desorption of nitrogen and oxygen on solid sorbents. In
partial pressure and μ is chemical potential. CO2 separation by polymeric and
other words, nitrogen and oxygen are separated in accordance with the
microporous membranes involve transport of molecular oxygen and hence yield
difference in their binding energies to the sorbent material [119]. low selectivity but high permeability, while separation via ITMs involves lattice
Activated carbons and most zeolite surfaces bind oxygen more strongly diffusion of oxide ions and hence offer absolute selectivity but low
than nitrogen. This leads to preferential enrichment of the gas phase permeability.
with nitrogen while the oxygen populates and stays on the sorbent
surface, resulting in preferential separation. Zeolites contain atomic across the membrane. Basically, there are three major types of mem­
scale crystallographic features in their structure that are conducive to brane technologies with vastly different operating principles as illus­
seize exclusion and sieving as well as selective adsorption. A variety of trated in Fig. 19. In polymeric membranes, air separation is achieved by
other sorbents such as activated carbon, silica, and alumina are also used relative dissolution and diffusion of O2 and N2 through the structural
for air separation [120,121]. As the relative binding strengths of voids in between the polymer chains, while in microporous inorganic
adsorbed species on solid surfaces vary with gas pressure and temper­ membranes, which are usually made of ceramics, separation is achieved
ature, adsorption-based air separation at the commercial scale is by selective, or differential adsorption, desorption and pore diffusion of
generally achieved using two related technologies, namely, pressure O2 and N2. Composite membranes combine the advantages of polymers
swing adsorption (PSA) [122] and vacuum swing adsorption (VSA), or and solid sorbents to modify separation properties such as selectivity
temperature swing adsorption (TSA). They employ typically zeolites for [125,126].These two basic types of membranes and their composites
the sorbent material, and are mature and reliable technologies that have operate typically at temperatures close to ambient.
been commercially available for several decades. By contrast, membranes based on lattice transport of oxide ions (i.e.,
Unlike cryogenic separation that requires extreme refrigeration, ion- or oxygen transport membranes (ITM or OTM)) proposed several
adsorption-based separation typically operates at ambient temperature, decades ago [127] operate above 700 ◦ C and are made of highly doped
but has modest production capacities of <1500 Nm3 O2 per hour. mixed conducting ceramics typically of the perovskite family of oxides
Moreover, many adsorption-desorption cycles may be necessary to that offer selective transport of oxide ions via vacancy diffusion through
obtain the desired purity of about 90% to 94% O2. The plant usually its crystal lattice [128], and accomplishes absolute separation, i.e.,
employs two reactors operating in tandem. The exhaust gasses from the production of 100% pure oxygen from air [129].
pressurized stage 1 reactor is transferred to the unpressurized stage 2 Among the three membrane categories, only ITM offers the ability to
reactor until the pressure in the latter is increased to the desired level. provide absolute selectivity to produce 100% O2, while polymeric and
The sequence is reversed in the next cycle. microporous membranes can produce enriched oxygen with purities
>35%O2. Membranes in all three categories require pressurization of air
4.1.3. Membrane-based air separation on the feed side in order to maintain a concentration gradient across the
Membrane-based air separation provides compact and modular so­ membrane that drives transport and preferential separation.
lutions for oxygen production [115,123,124]. Its modularity allows easy
scale up and more economical use of plant real estate, but is more
expensive than cryogenic separation, with higher capital costs and lower 4.2. Oxy-combustion
production rates. All membranes achieve separation under the driving
force of a chemical potential or pressure gradient externally imposed Oxy-combustion [130–134] employs pure or enriched oxygen

15
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 20. Schematic illustrating the simplified process steps and flow diagram for oxy-combustion [135].

usually from an air separation unit and yields flame temperatures in the for oxy-combustion [131,138]. Depending partly on the purity of oxy­
vicinity of 2500 ◦ C in high temperature boiler designs, and around gen used, oxy-combustion of fossil fuels produces nearly capture ready
1600 ◦ C in low temperature boiler designs. A typical process flow dia­ CO2 emissions. Unlike air-based pulverized coal power generation, the
gram of oxy-combustion is schematically illustrated in Fig. 20. The ad­ product stream from oxy-combustion usually contains about 70% CO2
vantages of oxy-combustion include retrofitting on existing power plants plus water vapor, as nitrogen does not enter the process stream,
[136], smaller physical size, and most importantly, the ability to pro­ rendering it easy to capture the CO2 from the flue stream. Absence of
duce capture-ready CO2 without the need for extensive post-separation nitrogen during oxy-combustion also produces nearly 75% less flue gas
operations, provided that highly pure oxygen is utilized for combustion. volume to process. When compared to air-based combustion,
On the other hand, the high flame temperatures in the boiler require oxy-combustion is reported to produce 66% less NOx and 30% less SOx
high temperature tolerant materials that are usually more expensive and emissions [139]. But due to the absence of the diluent N2, the concen­
with shorter lifetimes. This problem is mitigated by recycling part of the tration of impurities in the flue gas including inert gasses, HCl, SO2, ash,
combustion products back into the boiler in order to dilute the reaction and Hg are generally higher in oxy-combustion than in air-based com­
system and lower the boiler temperature as well as help reduce tem­ bustion processes, and their removal pose challenges [140,141]. In
perature gradients in the boiler. The advantage of oxy-combustion for particular, high concentrations of SO2, HCl as well as fly ash in the boiler
ease of carbon capture from fossil fuel-based power generation is illus­ lead to corrosion and require expensive materials of construction, and
trated in Table 5, which compares the compositions of boiler flue gas also increase the energy demand for both the CO2 compression and
streams from air-based and oxy-combustion of bituminous and transport steps towards storage purposes [136,142]. There have been
sub-bituminous coals [137]. excellent review articles that provide in-depth discussion and under­
Typically, cryogenic air separation is the preferred choice to produce standing of the benefits and challenges of oxy-fuel combustion [130,
highly enriched (~95%) oxygen for oxy-combustion [111,116,118]. 131,143,144].
However, other means of oxygen production have also been considered, Despite many advantages, oxy-combustion has not received the level
including polymeric membranes, composite membranes, nanotubular of interest and investments that it deserves. The major reason for this
carbon membranes, molecular sieve membranes, microporous ceramic was its cost. Indeed, the U.S. DOE had revived the former FutureGen
membranes as well as mixed ionically-electronically conducting (MIEC) program (until recently called, FutureGen 2.0 [145], which would have
dense ceramic membranes, also called ITMs [111]. Although ITMs offer provided about $1 billion assistance to FutureGen Alliance to develop
absolute selectivity for oxygen, slow and highly thermally activated and build an oxy-combustion power plant (instead of the originally
lattice transport of oxide ions in the MIEC ceramic structure limits their planned IGCC plant as designed in the former FutureGen program) in
oxygen production rates. In addition to their unique capability to pro­ place of the 168 MW coal-fired power plant near Meredosia, IL, in
duce 100% pure oxygen while excluding other gaseous species including cooperation with Ameren Energy Resources. This would have been the
N2, MIECs are also interesting for their catalytic properties and inves­ world’s first commercial-scale oxy-combustion power plant fully inte­
tigated for catalytic combustors that also provide an alternative option grated with carbon capture and permanent geologic storage. Unfortu­
nately, due to concerns about failing to fulfill the legislative deadlines
imposed on the industrial partners to meet the financial commitments
Table 5
Comparison of boiler flue gas compositions (dry basis) from air- and oxy- within the agreed time frame, DOE decided once again to pull its support
combustion of bituminous and sub-bituminous coals, highlighting the advan­ from the program in 2015. Nevertheless, there have been several pilot
tage of oxy-combustion in producing a highly concentrated CO2 flue stream for plant size demonstration projects in oxy-combustion many of them with
easy capture. 35% O2 in CO2 was employed for flue gas recycle in oxy-com­ capacities less than 100 MW with a handful of plants that are rated >
bustion [137]. 200 MW [132,145]. Although much of the challenges in CO2 capture
Flue Gas Bituminous Coal Sub-bituminous Coal and compression, flue gas treatment, and optimization and control of the
oxy-combustion process definitely benefit from decades-long experience
Air-fired Oxy-fired Air-fired Oxy-fired
(21% O2) (35%O2-CO2) (21% O2) (35%O2-CO2) in air-based combustion know-how, some of the remaining barriers that
need to be overcome include cost-effective and at-scale pure O2 supply,
CO2 (v%) 17.0 97.0 17.0 98.0
N2 Balance Balance Balance Balance
demonstration of high performance at large industrial scale, public
O2 (v%) 2.0 2.1 2.0 2.2 acceptance of CCS and appropriate pricing of carbon emissions.
SO2 (ppmv) 615 1431 175 372
NOx (ppmv) 583 1332 707 1183
CO (ppmv) 51 85 9 75 4.3. Chemical looping combustion

Chemical looping combustion [146–148] is an emerging technology


and involves delivering the oxygen for combustion in the form of a

16
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

suitable metal oxide, which serves as the oxygen carrier. The oxide is Table 6
reduced upon reaction with coal or natural gas in the fuel reactor. The Maximum oxygen carrying capacities of various metal oxide systems with and
reduced carrier is recycled into an air reactor where it is re-oxidized. without phase change during redox cycling.
Oxidation and reduction rates of the particular oxygen carrier Metal Oxide Oxygen Weight% Phase Change
employed are primarily governed by the defect chemistry and defect Cu/Cu2O 5.9 Yes
transport rates in the crystal lattice of the metal oxide [149]. Cu/CuO 11.2 Yes
The CLC process is illustrated in Fig. 21. The fuel reacts with the Ni/NiO 12.0 Yes
oxygen carrying oxide material in the fuel reactor, abstracting oxygen Co/CoO 11.9 Yes
Co/Co3O4 15.3 Yes
from the oxide and reducing it. As no nitrogen enters the reactor, the flue
Fe/Fe2O3 17.7 Yes
gasses from the reactor are highly concentrated in CO2 and H2O (steam). Fe/Fe3O4 16.0 Yes
The reduced oxide is regenerated in the air reactor and fed back into the Mn/Mn2O3 17.9 Yes
combustion reactor to repeat the CLC cycle. The combustion product Mn/MnO2 22.6 Yes
CO2 can easily be captured from the flue stream, primarily by La0.8Sr0.2Co0.5Fe0.5O3-x (x = 0.25) 0.95 No

condensing out the water vapor. Both reactors operate at temperatures


around 900 ◦ C. This is considerably lower than for oxy-combustion, and sites in the cation sublattice is charge-compensated by the creation of
provides operational and cost advantage. The efficiencies for power vacant sites in the oxygen sublattice in the crystal structure. Under a
generation and CO2 capture from coal-CLC were estimated to be 48% chemical potential gradient, these vacancies facilitate oxide ions to
and 98%, respectively [151]. diffuse through the crystal lattice, and provide absolute selectivity of
Unlike oxy-combustion, CLC does not require prior separation of oxygen transport. Doped perovskites such as A-site doped
oxygen from air. Instead it uses an internal redox cycle of a suitable La1–xSrxMnO3–d or A-and B-site doped La1–xSrxCo1–yFeyO3–d exhibit wide
metal-oxygen system to provide the oxygen for combustion. Two general nonstoichiometry in the oxygen sublattice and fast oxide ion transport
categories of oxygen carrying oxides have been investigated for CLC, [127,155], and allow rapid oxygen release or uptake without undergo­
namely, those undergo phase change during the redox cycle, and those ing structural or phase transformations. Such nonstoichiometric oxides
do not. Comparison of the theoretical oxygen carrying potentials of can be considered as alternative options for oxygen carriers provided of
select oxides is illustrated in Table 6. Major advantages of CLC include course that they possess the proper redox stability window within the
elimination of the capital, operating and energy costs associated with reduction and oxidation limits at the operating temperature during CLC
oxygen separation. On the other hand, CLC requires two reactors oper­ cycling [156]. On the other hand, these multicomponent non­
ating in tandem for closing the redox cycle, and involves high operating stoichiometric oxides are made of less abundant and costly elements and
temperatures. have heavy molar masses. Hence, simple oxide systems including Cu and
Oxygen carriers can be chosen from simple binary oxides based upon Ni, which are cheaper and more abundant have been the preferred ox­
common transition metals such as Cu, Ni, Fe, Mn in bulk particulate ygen carrier systems studied for CLC [152,153]. A recent review article
form or dispersed on a suitable support [152,153]. Naturally occurring discusses the successful testing of low cost oxygen carriers during
ores of Cu, Fe, Mn and Ca [154] as well as others, such as ilmenite [154] extended CLC cycling of solid fuels with encouraging results in carrier
have also been considered as oxygen carriers. As expected, oxidation lifetime, reactivity and performance, suggesting the viability of CLC for
and reduction cycles of these oxides are accompanied by significant cost effective CO2 capture [157].
changes in their molar volumes due to phase transformations. As a
result, internal stresses induced by these repetitive structural and phase
transformations in the material compromise and degrade their me­ 4.4. Integrated gasification combined cycle (IGCC)
chanical stability and durability upon prolonged use.
Alternatively, heavily defective multi-component oxides offer oxy­ Gasification offers many advantages over conventional combustion
gen abstraction and uptake without commensurate structural change, technologies, and there have been a handful of integrated gasification
which eliminates some of the durability problems of phase-change ox­ combined cycle (IGCC) plants many of which have been operating since
ygen carriers. On the other hand, their maximum oxygen carrying ca­ the mid 1990′ s in various countries around the world. More recent
pacities are inferior when compared to phase change oxide systems, as projects in the world include the 618 MW IGCC power plant in
indicated in Table 6. These oxides are typically of the parent ABO3 Edwardsport, IN, in the U.S.A., which has been operating since 2013, the
perovskite structure, where heavy extrinsic doping of the A and/or B 650 MW GreenGen IGCC polygeneration project in Tianjin City in China,
and 250 MW IGCC Nakoso Power Station in Fukushima, Japan.
Although they all aim at capturing at least part of their CO2 emissions,
none have demonstrated CCS at commercial scale yet. Recent cancel­
lations due to the widely underestimated budget of nearly AUD 7 billion
for the ZeroGen IGCC project in Australia, and massive cost overruns
($2.4 billion initial budget versus projected costs exceeding $7 billion)
of the Kemper County IGCC project [158] in Mississippi in the U.S.A at a
planned capacity of 582 MW with annual CO2 capture capacity of 3 Mt
was indicative of the treacherous and expensive pathway towards
achieving ‘clean coal power’ [159]. However, a recent commentary
from the International Energy Agency (IEA) [160] argued that such
failures should not discourage or deter the efforts to advance carbon
capture projects, which are essential for continued use of fossil fuels for
electric power production.
The advantages of IGCC include ease of capturing CO2 before it is
released into the environment, options for gas cleanup, and the flexi­
bility and versatility it offers for product distribution. The process pri­
marily involves the conversion of the solid fuel such as coal or biomass in
Fig. 21. Schematic flow diagram for chemical looping combustion of the presence of steam or carbon dioxide to a synthesis gas (or, syngas)
coal [150]. that consists primarily of CO and H2. In the context of natural gas,

17
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

similar reaction of steam with the primary component methane as well generation compared to most thermal conversion processes. This
as other hydrocarbons in natural gas is usually called steam reforming. approach, schematically illustrated in Figs. 23, involves a fuel cell -
However, gasification of solid fuels and reforming of gaseous (or, liquid) typically a solid oxide fuel cell (SOFC) - coupled to the process flow
fuels all produce syngas as product, and both are highly endothermic downstream of the steam gasification and gas clean up steps. Coal syngas
reactions, where heat must be provided to the system externally in order from gasification is cleaned of sulfur and other contaminants before it is
to drive the gasification and reforming reactions thermally. Heat re­ fed into a SOFC for electrochemical power generation at elevated tem­
quirements of major reactions relevant to reforming and gasification are peratures in excess of 700 ◦ C. In IGFC, air separation, gasification and
summarized in Table 7. After gas clean up, the syngas may undergo gas clean up processes involve separate steps, which are all external to
water gas shift reaction with additional steam, to produce more the fuel cell.
hydrogen which is typically combusted in a gas turbine for electricity Early attempts that adopted variants of this approach were pursued
generation. The term gasification is commonly used when the feed is by Westinghouse Power Corp [163–165] (now Siemens Power Genera­
coal or other solid fuels such as biomass, while reforming is used when tion Corp.), Stanford Research Institute [166–168] (SRI) and others in
the feed is natural gas or other gaseous or liquid hydrocarbons. the 1960′ s and 1970′ s. Currently the US Department of Energy through
Accordingly, the former process is labeled IGCC, while the latter is called its Advanced Energy Systems - Solid Oxide Fuel Cell Program (formerly,
natural gas combined cycle (NGCC). the Solid State Energy Alliance (SECA) program) [169] is supporting the
Gasification (or, reforming) provides an effective and environmen­ development of an integrated gasification fuel cell (IGFC) technology in
tally friendly way of converting coal (or, natural gas) into electricity a catalytic gasifier-SOFC system.
with carbon capture. It can also be employed for co-generation or pol­ More importantly, incorporation of fuel cells into fossil fuel-based
ygeneration, i.e., production of electricity as well as hydrogen and other power generation as in IGFC and natural gas fuel cell (NGFC) offer
high value chemicals [161]. Rather than combusting, the fuel is broken improved plant efficiencies and competitive economics for cost of
up into its chemical constituents during the gasification process. The electricity (COE). Comparative estimates by the U.S. Department of
primary purpose of gasification is to transform coal or other solid fuels Energy [170] that also accounts for the energy and economic burden of
into cleaner fuels with great environmental benefits in the form of CCS on plant efficiency as well as COE for several advanced coal- and
significantly lower emissions of SOx, NOx, particulates, heavy metals and natural gas-based thermoelectric power technologies demonstrate that
other coal contaminants. Coal gasification produces a highly concen­ inclusion of fuel cells offer the potential to achieve higher performance
trated (up to 90%) CO2 product stream and largely eliminates the need and efficiency at less cost than conventional power technologies. Results
for post separation operations to capture the product CO2. of this DOE study are presented in Fig. 24, and promise the prospects of
A generalized process flow diagram for a polygeneration IGCC plant, efficient and cost-effective generation of zero-carbon electric power
which employs an ASU to produce oxygen and utilizes it for steam from fossil fuels while fully capturing the highly concentrated CO2
gasification, is depicted in Fig. 22. It illustrates the ability of poly­ product with considerable ease. However, this approach still requires an
generation to capture CO2 and also produce electricity, hydrogen and ASU, which typically reduces the overall plant efficiency and increases
chemicals. The plant stack gasses are primarily CO2 and steam, which capital costs. For example, a recent study estimated that cryogenic-based
are easy to separate for carbon capture. The process provides flexibility ASU may account for nearly 14% of the capital cost and can consume
to employ either gas turbines to produce electricity or high temperature 10–40% of the gross power output of an oxy-fuel combustion plant
fuel cells to electrochemically convert syngas (or, hydrogen) into elec­ [111]. The need for an ASU can completely be eliminated by direct
tricity. It also offers co-generation capability by allowing power and conversion of the fuel to electricity inside the fuel cell proper, where the
hydrogen production via membrane separation (usually made of Pd al­ oxidation products are highly concentrated CO2 and moisture.
loys) as well as downstream production of useful chemicals and fuels,
achievable subsequently via the Fischer-Tropsch synthesis process.
As in oxy-combustion, oxygen is a key component of the IGCC pro­ 4.6. Carbon fuel cells
cess and is used for burning a portion of the coal to supply the heat
necessary to drive the endothermic gasification reaction. Typically With the impending challenge of mitigating climate change and the
cryogenic type of ASU is employed for large-scale oxygen production at need to capture CO2 from fossil fuel-based power generation, there have
~95% purity. The energy penalty imposed by the air separation unit is been renewed interest and promising advances in the development of
usually around 10–12% of the generated power and naturally burdens specialized fuel cells that aim towards efficient conversion of solid
the overall plant conversion efficiency [113]. carbonaceous fuels such as coal to electricity with capture-ready CO2
production [171–174]. These types of fuel cells are interchangeably
called carbon fuel cells (CFC), direct carbon fuel cells (DCFC), or
4.5. Integrated gasification fuel cell (IGFC) carbon-air fuel cells (CAFC), and offer two critically important advan­
tages, namely, higher conversion efficiencies and concentrated CO2
Incorporation of a high temperature fuel cell into the IGCC process product streams. Needless to say, high efficiencies and low emissions are
stream, as in integrated gasification fuel cell (IGFC), offers many ad­ imperative for sustainability and climate change mitigation.
vantages including higher efficiency of electrochemical power In essence, carbon fuel cell (CFC) is an electrochemical engine that is

Table 7
Enthalpies and qualitative rates of important reactions for gasification and reforming.
Process Reaction ΔH 298K Reaction Rate

Methane - reforming CH4 + H2O = CO + 3H2 + 205.8 kJ/molCH4 Slower than oxidation
Coal - gasification C + H2O = CO + H2 + 131.1 kJ/molC Slower than oxidation
Full oxidation C + O2 = CO2 –393.1 kJ/molC Rapid
Partial oxidation C + ½ O2 = CO –110.5 kJ/molC Rapid
Boudouard C + CO2 = 2CO + 172.3 kJ/molC Slower than oxidation
Water gas shift CO + H2O = CO2 + H2 –41.1 kJ/molCO Rapid
Methanation
(Fischer-Tropsch) CO + 3H2 = CH4 + H2O –205.8 kJ/molCO Slow
Direct methanation C + 2H2 = CH4 –74.9 kJ/molC Slow

18
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 22. Schematic flow diagram of a generalized polygeneration IGCC plant using steam gasification, with fuel cell and co-generation options, that is capable of
employing a variety of carbonaceous fuels as feedstock including coal, biomass, petroleum, and waste, and producing electricity, hydrogen, and syngas, all with CO2
capture [162].

Fig. 23. Simplified schematic flow diagram for integrated gasification fuel cell
(IGFC), where ASU denotes an air separation unit, and WGS indicates the water-
gas shift reactor. When gasifier is replaced with a reformer, this flow diagram
represents a natural gas fuel cell (NGFC) system. The dotted lines indicate the
process flow through WGS when hydrogen is desired to feed to the fuel
cell stack.

capable of converting the chemical energy of solid carbonaceous fuels


into electricity in a single chamber. In a similar manner, solid oxide fuel
cells (SOFC) have been developed for electrochemical conversion of
natural gas to electricity [109], and Bloom Energy in California has
already commercialized their natural gas-fed SOFC power units. The
SOFC structures and corresponding dominant reactions for carbon and
methane conversion, respectively, are illustrated in Fig. 25. These high
temperature fuel cells do not require an ASU as oxygen is supplied by
selective transport through an ionically conducting dense solid elec­
trolyte, typically yttria-stabilized zirconia (YSZ), thus avoiding the en­
ergy penalty for oxygen separation suffered by IGFC, IGCC and
oxy-combustion technologies. Oxide ion transport through the electro­
lyte is accompanied by electron flow in the external circuit. During the
fuel conversion process, eight electrons are released to the external
circuit for each CH4 molecule oxidized at the anode, while four electrons
are released for each carbon atom consumed in the solid fuel. These
electrons travel through the external circuit from the anode towards the
cathode to reduce the molecular oxygen in the air to oxide ions, and at
the same time perform useful electrical work. More importantly, the Fig. 24. Benefits of incorporating a solid oxide fuel cell (SOFC) into conven­
major reaction products exiting the fuel cell are highly concentrated CO2 tional coal and natural gas power generation for improving system efficiencies
and steam, which are easy to separate for CO2 capture. and lowering the cost of electricity. SOTA: State of art; PC: Pulverized coal;
There are several types of CFCs labeled after the type of electrolyte IGCC: Integrated gasification combined cycle; IGFC: Integrated gasification fuel
cel; NGCC: natural gas combined cycle (NGCC): Natural gas fuel cell. Reprinted
they employ. Solid oxide-based fuel cells (SO-CFC) employ ionically
from [110] with permission from The Electrochemical Society and IOP Pub­
conducting ceramic electrolytes such as yttria-stabilized zirconia that
lishing Ltd.
selectively transport oxide ions (O2− ) via lattice diffusion of oxygen
vacancies. Since this transport is highly thermally activated, SO-CFCs
operate at elevated temperatures typically >800 ◦ C. Molten carbonate-
based fuel cells (MC-CFC) employ eutectic mixtures of alkali-metal

19
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 25. Conceptual schematics of solid oxide-based high temperature fuel


cells and the net reactions for methane (left) and solid carbon conversion
(right) into electricity inside the fuel cell proper, producing primarily CO2 and
steam, which are easy to separate for CO2 capture. While the CO shuttle
mechanism is illustrated for both cells, the reaction pathway for H2 shuttle is
omitted for simplicity in the right panel, where H2O in the exhaust stream is
due to the oxidation of chemically bound hydrogen present in coal and in
other solid carbonaceous fuels. Reprinted from [175] with permission from
The Electrochemical Society and IOP Publishing Ltd.

carbonates including Li2CO3, Na2CO3, and K2CO3 as the molten elec­ ions in the case of MCFC) through the fuel cell electrolyte facilitates the
trolyte that transports oxygen in the form of carbonate ions (CO32− ) and gasification and oxidation reactions of the fossil fuel inside the cell, and
operate typically at temperatures in the range 600 to 750 ◦ C depending eliminates the need for an ASU, which otherwise reduces the overall
upon the eutectic composition. There is also a hybrid-CFC version that plant efficiency typically by 8% to 12% [113,131]., Although significant
employs a solid oxide electrolyte in series with a molten carbonate advances were made in internally and thermally integrating gasification
electrolyte, and operates typically ~750 ◦ C. In all these CFC configura­ with solid oxide fuel cells [107,176–178], carbon conversion in fuel cells
tions, oxygen that is required for the oxidation reaction of the fuel is possesses inherent difficulties and technical challenges as examined
supplied across the electrolyte in ionic form through a thermodynami­ elsewhere [30,179–181] and plausible mechanistic pathways to achieve
cally downhill process. electrochemical conversion in a variety of CFC configurations were
CFCs achieve conversion of coal directly to electrical power via proposed [182]. Much progress has been made recently in both under­
electrochemical oxidation, and not by combustion. This is an important standing fundamental issues as well as improving cell performance, but
distinction. It should be noted, however, “direct” conversion does not much work still remains to advance this nascent and important
necessarily suggest direct electro-oxidation of the fossil fuel itself technology.
directly, as oxidation products CO2 and H2O undergo further gasifica­ Although gaseous (e.g., methane) as well as solid carbonaceous fuels
tion in the case of coal or other solid fuels, and reforming in the case of such as coal and biomass [22,23,30,107–109] can be converted to
natural gas, providing positive feedback reaction loops as shown in electrical energy in SO-CFCs with the emission of highly concentrated
Fig. 25. These gasification and reforming reactions further lead to water CO2 flue streams, there are technical challenges to be overcome. These
gas shift reactions. So in a strict sense, the term “direct” indicates that include the high energy barrier for methane oxidation due to the diffi­
conversion process takes place in the fuel cell proper, regardless of the culty in activating the strong C-H bond [109], difficulty to oxidize solid
details and the nature of the chemical and electrochemical reactions carbon fuels due to poor solid-to-solid contact between the fuel and the
taking place. anode surface [182], anode deactivation and cell degradation due to
Another major advantage of CFCs is to reduce the number of process sulfur poisoning, interplay among electrochemical oxidation of the fuel
steps and more importantly negate the need of oxygen separation in an with chemical reactions of gasification or reforming and water gas shift,
ASU. This is in contrast to gasification and combustion technologies that difficulty in achieving truly direct electrochemical oxidation of the fuel
require not only air separation units for oxygen production - for IGCC, [30], thermal and microstructural durability and stability of cell mate­
NGCC, IGFC and NGFC technologies – but also multiple process steps rials and components, and among others, materials and fabrication
thereafter (see Figs. 22 and 23). Collectively, these multitude of indi­ costs.
vidual process steps introduce efficiency losses as expected. By contrast, For MC-CFCs, the challenges are different. Molten carbonate fuel
CFCs advance the potential opportunities beyond IGFC and NGFC by cells cannot directly process or convert methane, as the solubility of
internally and thermally integrating the gasification (or, reforming, in methane in the alkali carbonate eutectic is rather small. However, they
the case of natural gas) process inside the fuel cell proper, or chamber. can employ methane in the form of hydrogen or syngas (i.e., CO-H2
This offers greater benefits in performance and efficiency, but more mixture) only after methane undergoes steam reforming in an external
importantly, ease of carbon capture, as CFCs eliminate the need for an reformer [30,109]. On the other hand, MC-CFCs are suitable to handle
external ASU and/or a separate gasifier. The oxidant needed to gasify and convert solid carbonaceous fuels such as coal, biomass, waste etc.
and oxidize the fuel is provided electrochemically through the fuel cell’s into electrical energy [176]. The challenges to conversion of solid fuels
electronically insulating but ionically conducting electrolyte. Modified in MC-CFCs include hot corrosion and stability of the molten electrolyte,
architectures of typical solid oxide- (SOFC) [107] and molten degree of percolation (or electrical connectivity) as well as wetting of
carbonate-based fuel cells (MCFC) [178[,or their hybrids, namely, solid fuel particles in the molten electrolyte, sensitivity to moisture – a
molten carbonate/solid oxide electrolyte configurations [177] have byproduct of coal or biomass oxidation, parasitic side reactions
been shown to be suitable configurations for achieving efficient con­ including the Boudouard reaction, and slower overall rates typical of
version of solid fuels to electricity. liquid media for all rate processes leading to limited cell performance
Obviously, combining multiple process steps of the IGFC and NGCC [30,179,181]. Hybrid-CFCs [180] borrow these shared challenges from
processes into a single process inside the fuel cell proper promises higher each, in addition to the long-term chemical stability of the solid oxide
electrical conversion efficiencies exceeding 60% [108], simplicity in ceramic electrolyte in the highly corrosive and strong solvation action of
plant design, better thermal management and control, smaller real es­ the molten carbonate electrolyte [30,177].
tate, and potentially lower capital and operational costs. It may be ex­
pected that these savings promise an even more favorable comparison 5. Post-combustion carbon capture technologies and materials
with other power generation technologies presented in Fig. 24. Selective
transport of the oxidant (oxide ions in the case of SOFC, and carbonate Post-combustion carbon-capture offers the most expedient near term

20
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Table 8 CO2 concentration of the emission source, as well as the cost and ease of
Typical gas compositions of the flue streams from coal-fired power plant, and deployment of the separation process. Many of these technological op­
water gas shifted coal gasification and steam methane reforming plants (HCs: tions and their challenges are presented in recent reports [78,89,
hydrocarbons). 184–190], and also discussed below.
Flue Gas Coal-Fired Shifted Coal Shifted Steam
Component Power Plant Gasification, vol% Reforming, vol%

N2 73 to 77 vol% 0.3 to 2.3 5.1. CO2 capture by liquid sorbents


CO2 15 to 16 vol% 25 to 35 15 to 25
CO 20 ppmv 0.5 to 0.7 1 to 3 Post-combustion CO2 capture in liquid solvents can be achieved both
CH4 0 3 to 6
by physical or chemical absorption. The latter usually involves a
H2O 5 to 7 vol% 15 to 40
H2 30 to 50 70 to 80 chemical reaction step, which is not the case in physical absorption.
O2 3 to 4 vol% Both types of absorption rely on the interfacial equilibrium between the
SO2 800 ppmv flue gas stream and the absorbent liquid phase, and employ a two-step
SO3 10 ppmv
process. First, the product flue gas is preferentially stripped off from
H2S 0.1 to 0.2
NOx 500 ppmv
its CO2 content by contact with a suitable solvent in an absorption tower,
HCl 100 ppmv then the absorbed CO2 is removed from the solvent in a regeneration
HCs 10 ppmv tower, which also recovers the solvent for the next batch of CO2 capture
Hg 0.001 ppmv step. As the solubility of gasses in liquids increases with decreasing
Ar 0.04
temperature and increased pressure, the absorption process is usually
carried out at low temperature and/or high pressure, while the regen­
eration step to recover both the CO2 and the solvent involve high tem­
peratures and/or low pressures. Moreover, the relatively low
concentrations of CO2 in the flue gas from natural gas turbines (3% to
potential for mitigating carbon emissions especially from point sources, 5%) and coal-fired plants (10% to 15%) require the use of absorption
because many of the technologies in this category can be added to towers with large contactor areas and consequently higher capital and
existing power plants without major disruptions or changes in the pro­ energy costs for effective CO2 capture.
cess flow. Post-combustion capture technologies are designed to aim at Historically, liquid solvent processes primarily based on aqueous
separating and capturing CO2 from N2-rich product gasses after the air- solutions of alkali metal hydroxides that chemically react and remove
combustion step but before releasing the flue gasses into the atmo­ CO2 have been employed for nearly a century as part of the life support
sphere, although other contaminants may also impact the choice of systems in confined environments including submarines and space
capturing method. Typical composition of flue gas streams from coal- crafts, where the resulting precipitates of the insoluble alkali metal
fired power plants are compared in Table 8 to flue gas compositions carbonates are filtered and removed from the solution. This specialized
from water gas shifted coal gasification and water gas shifted steam process, however, is not suitable for carbon capture from large point
reforming of methane. With practically no N2 present in the product flue sources such as coal-fired power plants.
stream, it is clear that pre-combustion does not necessitate separation of For effective capture of CO2 from massive volumes of the flue gasses
CO2 in the latter two processes. For post-combustion from coal-fired emitted from power plants, exceptionally large interfacial contact area is
power generation, on the other hand, the method of choice demands required to overcome the mass transfer resistance that exists at the gas/
high selectivity for CO2 in the presence of high concentrations of N2, liquid interface of the absorber column. This mass transfer limitation
with minimal interference from other contaminants. generally necessitates the use of large volume absorbers, which add
A wide variety of CO2 capture technologies are employed down­ significant cost to the capture system and may constitute up to 55% of
stream of the combustion process. Mechanistically, they differ widely in the total capital cost [191].
the way each accomplishes CO2 separation and capture. They primarily
involve absorption, adsorption, mineralization, membrane and cryo­ 5.1.1. Capture by chemical absorption
genic separation, as well as electrochemical separation via fuel cells. Chemical absorption involves a reaction step between the solvent
Regardless of the type, however, CO2 separation and capture adds and the solute, in this case CO2. Generally amine-based solutions are
considerable penalty to the overall process in both energy and cost. considered for carbon capture as they have been used already for gas
Indeed, a recent thermodynamic study [106] for post-combustion clean up in several industries, although the processes are costly and
capture and storage of CO2 from coal-fired plants has estimated the energy intensive. To increase capture rate and efficiency as well as to
lower bound of the energy penalty to be 24 kJ/molCO2 into a 2-km deep lower costs, numerous compounds are being investigated, including
underground injection cavity, assuming perfect 2nd law efficiencies. ammonium-based solvents.
Moreover, the energy requirements for carbon capture and storage
(CCS) for pressure swing adsorption (PSA) and temperature swing 5.1.1.1. CO2 capture in amine-based solutions. Aqueous amine solutions
adsorption (TSA) processes were estimated to be 75 kJ/molCO2 and have been employed commercially in the gas industry for more than half
45 kJ/molCO2, respectively, assuming no compression losses and per­ a century to strip CO2 by chemical absorption with CO2 recovery effi­
fect separation. Analyzing the spread of the actual thermal efficiencies of ciencies up to 98% [192], although for large scale applications 70% to
420 large size U.S. installed coal-fired plants, which showed a near 95% capture efficiency is likely more realistic. Primary challenges for
Gaussian distribution of end efficiencies between 46% and 18% where CO2 capture in aqueous amine solutions include corrosion to equipment
the average was ~33%, this study further concluded that the energy and materials, thermal degradation of amines, and high energy con­
penalty for capturing and storing 80% of the CO2 emissions from the sumption. The production of amines primarily involves the reaction of
entire U.S. pulverized coal (PC) fired power fleet is expected to require ammonia with ethylene oxide. However, commercial production of
an additional 390 to 600 million tons of coal (at nominal 25 MJ/kg), or ammonia is highly energy intensive (consumes >1% world’s energy
an additional 69 to 92 GW of CO2-free base-load of power, which con­ supply) and inherently has a large CO2 footprint on its own. The esti­
stitutes 15% to 20% reduction in the overall electricity production. mates of CO2 emissions from ammonia synthesis range from 1.15 to 1.40
Many technological options exist for separating CO2 from flue gas kgCO2/kgNH3 [193] to 1.87 kgCO2/kgNH3 [194]. Hence, the cumula­
streams. Selection of the optimal process largely depends on factors such tive impact of amine-based CO2 capture on the environmental is quite
as the type, production volume (or, rate), flue gas composition, and the significant beyond the capture process per se, and must also account for

21
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

the CO2 emitted during the upstream production of ammonia. Using a feed gas of 15vol%CO2 and 85vol% N2 at 40 ◦ C, a recent
The tail-end technology of amine-based separation is reasonably study on the thermodynamic and kinetic characterization of nearly 30
mature, and is readily suitable and adoptable for CO2 capture from the different aqueous 30% amine solutions at 0.1 MPa pressure found that
flue streams of coal-fired power plants and other point emission sources hexamethylenediamine (HMD) exhibited the highest loading of 1.35
[195–197]. The process involves CO2 absorption from a flue gas stream molCO2/mol of amine, and triethanolamine (TEA) showed the lowest
into a low volatility aqueous amine solution near ambient conditions. loading at 0.39 molCO2/mole amine, while the loading levels of the
Upon absorption in the scrubber, the amine solution is then regenerated commonly employed monoethanolamine (MEA) and diethanolamine
by stripping it with steam at 100 to 120 ◦ C. The resulting vapor is (DEA) were 0.58 and 0.53 mol CO2/mole of amine, respectively [198].
condensed to remove water and the released CO2 is recovered and The underlying principle of amine-based absorption involves the
generally compressed to ~15 MPa for storage. Typically alkolamines are formation of water-soluble compounds such as carbamates and bi­
employed for this purpose and the most commonly used amine solutions carbonates when amines react with CO2. This reaction is driven by the
contain monoethanolamine (MEA), diethanolamine (DEA), or N-meth­ fact that amine solutions are basic in character while CO2 is mildly
yldiethynolamine (MDEA). In practice, 20% to 30% MEA solutions are acidic, thus facilitating a chemical reaction.
more commonly employed for CO2 capture than others. Other pro­ The absorption reactions for the primary and secondary amines with
spective solvents of this group also include piperazine (PZ) and its de­ CO2 form a zwitterion (a neutral molecule containing negatively and
rivatives, which exhibit better thermal stability and higher CO2 capture positively charged groups) first and then a carbamate salt as expressed
capacity than MEA, but also have higher vapor pressure that may limit by,
their practical application. At the same time, greater capacity for CO2
uptake minimizes heat losses during heating and cooling between ab­ RR’NH + CO2 = RR’NH+COO− (zwitterion) (2)
sorption and regeneration cycles. These advanced amine systems can + −
RR’NH + RR’NH COO = RR’NCOO (carbamate) + RR’NH − +
(3)
capture CO2 with a heat duty of < 2.7 MJ/tCO2 and work <
250 kWh/tCO2 including compression of the captured CO2 to 15 MPa The reaction with the tertiary amine forms bicarbonate but not carba­
[80]. Many of the physicochemical properties of select amine solutions mate. The overall reaction is given by,
are summarized in Table 9. They exhibit CO2 absorption fluxes that vary
2RR’NH + CO2 = RR’NCOO− + RR’NH2+ (4)
between 2.4 × 10− 7 molCO2/m2.Pa (for AMP) and 8.6 × 10− 7 mol­
CO2/m2.Pa (for AMP/PZ). Naturally high capacities for CO2 uptake as −
RR’NCOO + H2O = RR’NH + HCO3 −
(5)
well as high temperature limitations for amine regeneration are desired
properties, although high Pmax is indicative of high vapor pressure and is The major drawback for amine-scrubbing is the high energy
undesirable. Some piperazine (PZ)-based solutions exhibit relatively requirement for the solvent regeneration step at high temperatures.
high CO2 uptake capacities approaching 1 molCO2/kg with sufficiently Especially MEA is not a particularly energy efficient and environmen­
high tolerance to high temperatures during regeneration when tally friendly solvent for CO2 capture. It is also a costly process. An early
compared to commonly employed monoethanolamine (MEA) solutions. techno-economic analysis [199] of MEA-based CO2 capture from a
General categorization of alkolamines includes primary, secondary coal-fired power plant of 500 MW gross capacity, the energy penalty for
and tertiary amines containing at least one hydroxyl and amine group. capturing 84% of the CO2 brings the net plant capacity from 462 MW for
Examples of primary amines are monoethanolamine (MEA) and digly­ the control case without CO2 capture down to 326 MW with capture, and
colamine, while diethanolamine (DEA) and diisopropanolamine are increases the cost of electricity nearly 100% with capture. It clearly
examples of secondary amines. N-methyldiethynolamine (MDEA) and shows that CO2 capture demands a parasitic power of more than 130
triethanolamine (TEA) are tertiary amines. The reactivity towards CO2 MW or 32% of the produced power capacity. This techno-economic
increases with amine ranking, while the CO2 capacity does not. For study also concluded that the cost of carbon capture is expected to be
example, the capacity is about 1 molCO2/mol of tertiary amine, which is higher for retrofitting old power plants with carbon capture capability
significantly higher than the average CO2 capacity of ~0.5 molCO2/mol than in newly designed plants.
of primary and secondary amine. On the other hand, CO2 reactivity A later study summarized the results of two earlier techno-economic
follows rank and the uptake reaction rate constants for CO2 in primary, analyses on the same 450 MW power plant conducted in 2001 and 2006
secondary and tertiary amines at 25 ◦ C are respectively, 7000 m3/s.kmol for two different MEA-scrubbing solutions used for CO2 removal [195].
for MEA, 1200 m3/s.kmol for DEA, and 3.5 m3/s.kmol for MDEA [184]. The results show that the concentration of the MEA solutions used for

Table 9
Comparative physicochemical properties of select amine solutions, where capacity is per kg of water+amine solution, Tmax is the upper temperature limit for the onset
of thermal degradation, and Pmax is the maximum boiler pressure during regeneration. Adapted from [80].
Amine Molality Capacity –ΔHabs Tmax Pmax
(m) (molCO2/kg) (kJ/mol) (◦ C) (bar)

Monoethanolamine (MEA) 11 0.66 70 125 2.7


MEA 7 0.47 70 121 2.2
Piperazine (PZ) 8 0.79 64 163 14.3
MEA/PZ 7/2 0.62 80 104 0.7
Methyldiethanolamine (MDEA)/PZ 5/5 0.99 70 120 1.8
MDEA/PZ 7/2 0.8 68 120 1.4
2-Amino-2-methyl propanol (AMP) 5 0.96 73 140 6.1
AMP/PZ 2.3/5 0.7 71 134 4.5
2-Piperidine ethanol 8 1.23 73 127 3.3
2-Methyl piperazine 8 0.93 72 151 9.9
2-Methyl piperazine/PZ 4/4 0.84 70 155 10.3
Kglycinate 6 0.35 69 120 1.08

22
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

capture has a significant impact on both the cost and energy penalty, ammonium chloride with lime.
such that the overall cost of CO2 capture is estimated to decrease more
than 32% going from 20% MEA solution to 30% MEA solution, while the 2NH4Cl + Ca(OH)2 = 2NH3 + CaCl2 + 2H2O (7)
energy penalty for capture decreases 27%, respectively. Lime is produced from limestone by the calcination reaction followed by
Many of the amine solutions also pose technical challenges, hydration,
including relatively low solubilities and hence capacities for CO2, which
necessitate the use of large absorbers and hence high cost. In addition, CaCO3 = CaO + CO2 (8)
amines that are vaporized from the solution or emitted from post-
CaO + H2O = Ca(OH)2 (9)
combustion carbon capture operation may undergo photo-oxidation in
the atmosphere [200], where the products of oxidative and thermal Thus the overall reaction for the dual-alkali process is,
degradation of amines pose further problems for the environment [201,
202]. Particularly, many of the degradation products such as nitrosa­ CO2 + 2NaCl + CaCO3 + H2O = 2NaHCO3 + CaCl2 (10)
mines and nitramines that may form by reaction of amines with SO2, In this dual-alkali process, ammonium serves as the primary alkali
NO2 and O2 in the flue gas are known carcinogens, and pose potential and lime as the secondary alkali. The precipitated sodium bicarbonate
harm to human health. Degradation products also lead to corrosion can be collected, and then heated to form sodium carbonate (Na2CO3).
problems in the process equipment and piping. Advanced amines like Despite its simplicity and low cost for ammonia, this process has also
piperazine and ethyldiethanolamine are chemically more stable and drawbacks. It is energy intensive, partly due to the heat requirement for
more resistant to degradation than conventional amines, but they are calcination of the limestone, which serves as the secondary alkali. Also,
also more expensive, and more vulnerable to interactions with other flue for every two moles of CO2 removed from the flue stream, one mole of
gas components such as SO2, NOx and fly ash, requiring pretreatment to CO2 is released back to the atmosphere during the calcination of
remove these contaminants from the flue gas prior to the absorption limestone.
step. Similar to amines, ammonia based CO2 capture also has two
Mixtures of amines in polyethylene glycol (PEG) are also explored important drawbacks, one due to huge energy consumption, and the
[203]. PEG has low vapor pressure and almost half the heat capacity of other due to its large carbon footprint, both of which has to do with the
water, which leads to lower energy penalty during regeneration and upstream production of ammonia in the first place. Ammonia, as one of
significantly less corrosion. However, most of these mixtures have Tmax the largest quantity chemicals in the world (230 million metric tons in
values around 90 ◦ C that lead to onset of undesirable weight loss at 2018), is produced commercially via the Haber-Bosch process that in­
temperatures significantly lower than other amine solutions (see volves steam reforming of methane to produce the hydrogen needed to
Table 9), and pose challenges for CO2 capture at higher temperatures. react with nitrogen under high pressure (typically 10 to 25 MPa) and
Alternative pathways to reduce the energy burden of the regeneration elevated temperatures (350 to 550 ◦ C) over a Fe-based catalyst. This
step include augmentation with solar thermal energy integrated into the process not only consumes nearly 1% of the world’s energy, but also
CO2 capture process [204], and replacing thermal regeneration with produces copious amounts of emissions up to 1.87 kgCO2/kgNH3 [194].
capacitive deionization (CDI), where charged dissolved ions in the Furthermore, it is estimated that nearly 40% of the fertilizer nitrogen,
amine solution depicted in Eqn. (2) through Eqn. (5) are collected after which is directly derived from ammonia, is lost to the environment and
electrosorption on the anode and cathode surfaces by means of an into the atmosphere - as dinitrogen and NOx [209]. This loss of fertilizer
externally imposed potential, and the regenerated ion-free amine solu­ represents not only a significant waste of energy, but also adverse
tion is then fed back to the CO2 absorber, resulting in overall energy environmental impact of the reactive nitrogen species in the atmosphere
savings of 10% to 45% [205]. as well as in the waterways seeping into the food supply chain. On the up
side, the byproducts of ammonia absorption typically include NH4CO3,
5.1.1.2. CO2 capture in ammonia-based solutions. In an attempt to NH4NO3, and (NH4)2SO4, which can all be used as fertilizers. These
overcome or bypass some of the technical drawbacks and deficiencies of upstream and downstream considerations are important for compre­
amine-based scrubbing, absorbents other than amines have also been hensive assessment and selection of the suitable CO2 capture
investigated for CO2 capture. One of these is the “chilled” ammonia technology.
process, where the flue gas reacts with the aqueous ammonia solution in Due to their non-volatility, low corrosivity, and non-toxicity, sodium
a wet scrubber to form ammonium bicarbonate (NH4HCO3) or ammo­ carbonate-bicarbonate slurries were also explored for CO2 capture from
nium carbonate ((NH4)2CO3). Although the regeneration of the coal-fired power plants and the energy requirement for regeneration was
ammonia solution requires heat input in order to thermally decompose estimated to be 3.22 MJ/kgCO2 as opposed to a higher energy penalty of
ammonium bicarbonate or carbonate, it was reported that the aqueous 3.8 MJ/kgCO2 for the MEA-based capture process [210]. Although
ammonia process is significantly more energy efficient and requires up environmentally friendly, sodium carbonate-based capture exhibit low
to 60% less energy than the MEA scrubbing process [206]. Likewise, the absorption rates compared to many of the amine-based solutions, and
total energy requirement for ammonia-based CO2 capture was estimated require larger scrubbers with greater contact areas.
to be 1147 kJ/kgCO2 compared to a much higher penalty of
4215 kJ/kgCO2 for a MEA-based capture process using 30wt% MEA 5.2.1. Capture by physical absorption
solution [207]. Prior to scrubbing, the flue gas undergoes pretreatment Capture by physical absorption does not involve chemical reaction
in order to oxidize the NO and SO2 impurities to NO2 and SO3, respec­ but instead dissolution of CO2 in the solvent. In principle, the limitation
tively. Correspondingly, the other by-products of scrubbing are usually for physical absorption of any gas in a liquid is governed by Henry’s law,
ammonium sulfate ((NH4)2SO4) and nitrate ((NH4NO3), which are which simply states that at constant temperature, the concentration, ci,
valuable chemicals used also as fertilizers. of a particular gas, i, dissolved in a unit volume of liquid is related to its
Similarly, a modified Solvay process was proposed [208] that em­ partial pressure, pi, in equilibrium with that liquid. At infinite dilution,
ploys a dual-alkali approach using ammonia to catalyze the CO2 reaction this can be expressed by
with sodium chloride to form and precipitate sodium bicarbonate by the
reaction, Pi = kH .Ci (11)

CO2 + NaCl + NH3 + H2O = NaHCO3 + NH4Cl (6) Here, kH represents Henry’s constant. For CO2 dissolved in water at
room temperature, kH = 3.44 mol/m3 MPa. It is clear from Eqn. (11) that
During the regeneration step, ammonia is recovered by reacting the capacity of the solvent is limited by the CO2 partial pressure, and

23
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Table 10 investigated for their CO2 solubility, stability and uptake rates over a
Commercial liquid sorbents for CO2 capture by physical absorption [184]. wide pressure and temperature regime. The compounds included per­
Process Trade Name Primary Component fluoroperhydrofluorene (C14F24), perfluoroperhydrophenanthrene
(C17F30), and perfluorocyclomethyldecalin (C13F22), known respectively
Selexsol Dimethyl ether (or, propylene glycol)
Rectisol Methanol as PP10, PP11, and PP25. The results indicated that CO2 is up to 7 times
Purisol N-methylpyrrolidone more soluble than N2 in all the three PFCs tested, and found them to be
Morphysorb Morpholine stable and reliable solvents for selective capture of CO2 from flue
Fluor Propylene carbonate streams at temperatures up to 227 ◦ C and pressures up to 3 MPa.
Nevertheless, solvent loss by evaporation, relatively low mass transfer
coefficients for the gasses in PFCs, as well as the decrease in the gas/­
liquid interfacial area with increasing viscosity pose technical chal­
lenges for the use of PFCs for effective carbon capture.
thus higher pressures favor higher dissolved gas content. This poses a
challenge for CO2 capture from the flue gas streams of fossil fuel power 5.2.1.2. Capture in ionic liquids. Ionic liquids (IL) are considered viable
plants, where the concentration of CO2 is rather low, typically 10% to alternatives to amine-based solvents for CO2 capture, and generally
15% for coal-fired and 3% to 5% for natural gas power plants. On the exhibit exceptional chemical and thermal stability, negligibly low vapor
other hand, CO2 capture by physical absorption from flue streams of pressure, non-flammability, recyclability, high polarity, high viscosity,
IGCC plants looks more attractive due to the significantly higher con­ and low toxicity [211–213]. For these reasons, ionic liquids have gained
centrations of CO2. interest in the last two decades for a wide range of applications. These
As the solubility of gasses in liquids usually decreases with increasing include solvents for gas separation, electrolytes for batteries, super­
temperature, effective separation of CO2 by physical absorption in liquid capacitors and other electrochemical devices [214], or inert solvents for
solvents involves operation at reduced temperatures and high pressures catalysis [215].
in order to achieve higher concentrations of dissolved CO2. This is fol­ Ionic liquids constitute an extremely rich class of high molecular
lowed by a desorption step at reduced pressure and/or increased tem­ weight organic salts most of which have melting points near room
perature, where CO2 is released and collected from the liquid solvent. temperature and contain an organic cation, and either an inorganic or an
Several commercial processes based on physical absorption are available organic anion. Some of the more commonly employed anions and cat­
for industrial scale separation and capture of CO2 and other impurities ions are given in Fig. 26. The number of possible IL compositions is quite
from flue streams as tabulated in Table 10. Selexsol process is preferred large, and the properties of ILs largely depend upon the choice of anions
for the removal acid gasses (CO2 and H2S), and offers several advantages and cations. So tuning the properties of ILs to obtain enhanced CO2
including low vapor pressure, low toxicity and low corrosion activity.
Similarly, the primary solvent methanol for Rectisol process is stable and
less corrosive. Both Selexsol and Rectisol are commercially employed
not only for CO2 separation but also for H2S capture from flue streams.
However, both are energy intensive due to the thermal requirements of
the absorption and desorption steps. By contrast, Purisol process offers
lower energy consumption, while Morphysorb process offers 30% to
40% reduction in operating costs compared to Selexsol [184]. The high
solubility of CO2 in polypropylene carbonate renders the Fluor process
attractive for capture, while the weaker bonding between the solvent
and CO2 allows easy separation at reduced pressure [186].
Although CO2 capture via physical absorption processes generally
requires less energy for CO2 release and solvent regeneration steps, they
are severely limited by uptake capacity and their operating temperature
regime. Cooling the flue gas stream down to ambient temperature is
generally required prior to the absorption step to capture CO2 in the
solvent. This adversely impacts the thermal efficiency of the process. For
these and other reasons including the lack of a regulatory or financial
penalty (or, price) on carbon emissions, physical absorption technolo­
gies have not yet received wide scale attention for CO2 capture at the
industrial scale. There is need for developing better and more stable
solvents with higher solubility for CO2 at higher temperatures than those
available today.

5.2.1.1. Capture in fluorinated solvents. Perfluorinated compounds


(PFC) offer high stability due to the strong C-F bonds in their molecular
structure, and consequently, have moderate vapor pressures and high
boiling points. They also exhibit low molecular interactions due to the
repulsive nature of the fluorine atoms in the structure and have negli­
gibly small dipole moments. These attributes provide high solubility for
CO2 during absorption, as well as a low energy barrier to release the CO2
during the solvent regeneration step at reduced pressures or increased
Fig. 26. Examples of types of anions and cations used commonly in ionic liq­
temperatures.
uids, where Ri’s are often alkyl groups. Reprinted from [216] with permission
In an earlier study [185] reported by the National Energy Technol­
from the American Chemical Society.
ogy Laboratory (NETL) of the U.S. Department of Energy (DOE) for the
removal of CO2 by wet scrubbing of flue gasses from large point sources
such as power plants, several different types of PFCs have been

24
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

solubility for effective capture is an important technical challenge, but at [216]. Better understanding of the dominant dissolution mechanisms
the same time an attractive opportunity for CO2 separation by absorp­ requires further investigation of this interesting class of organic salts.
tion in ILs [185,217], and by IL-supported and facilitated transport Among the vast variety of IL compositions, imidazolium-based ionic
membranes [218]. liquids attracted arguably the most attention due to their remarkably
Results of experimental work using various spectroscopies as well as high solubility for CO2 [216,217,221]. The solubility of CO2 at 25 ◦ C in
theoretical studies using molecular simulations are in good general several imidazolium-based ionic liquid compositions is presented in
agreement that interactions between CO2 and the anion moiety of the Fig. 27. The observed difference between hydrogen- and
ionic compound dominate the extent of solubility of CO2 in ILs, while the methyl-substituted ILs in the temperature range 10 to 50 ◦ C suggests this
cation plays more of a secondary role [216,217]. This interaction is substitution may influence the CO2-cation interactions. For all temper­
described as of Lewis acid-base type, where the anion serves as the Lewis atures tested, ionic liquids with the [Tf2N] anion exhibited the highest
base and CO2 acts as the Lewis acid. Longer cation alkyl side chain CO2 solubility [217]. Similarly, ionic liquids based on the cation trie­
lengths seem to increase CO2 solubility by decreasing the extent of thylenetetramine (TETA) with suitable anions, such as with tetra­
cation-anion interactions and possibly increasing the free volume for floroborate, i.e., (TETAH)+(BF4)− exhibits a CO2 absorption capacity of
CO2 in the solvent. Furthermore, there is evidence that, especially in the 0.96 molCO2/molIL at room temperature while addition of 40% water
case of imidazolium-based ILs, no significant difference exists for CO2 improved the capacity to 2.04 molCO2/molIL [222]. Similarly, absorp­
solubility in ILs with hydrogen- versus methyl-terminated anion, sug­ tion capacity is shown to relate closely to the ionic structure, and for the
gesting that CO2-anion interactions cannot alone explain CO2 solubility same cation has followed the order (NO3) > (BF4) > (SO4), with 2
in ILs, and other factors may be in play [217]. These secondary factors (TETAH)+(NO3)2− containing 40% water exhibited the best CO2 ca­
may involve the free volume mechanism, where CO2 molecules reside in pacity [223]. However, these imidazolium- and amine-based ionic liq­
the free spaces or cavities in liquid, and also entropic effects, and mo­ uids also have ammonia as the starting chemical in their production, and
lecular interactions such as electrostatic, hydrogen bonding and van der suffer from similar disadvantages in life cycle considerations regarding
Waals forces between the species in the liquid phase. Computational high energy penalties and large carbon footprints of ammonia, as
studies also indicated that attractive van der Waals forces dominate the pointed out in the previous sections of this article on CO2 capture by
solubility of CO2, and electrostatic interactions has a secondary influ­ chemical absorption.
ence on the excess enthalpy of dissolution while hydrogen bonding did For all gas separation applications, selectivity is as important a
not seem to have a significant impact [219]. Based on these findings, this consideration as solubility. This is particularly true for CO2 capture from
study further investigated the impact of increased van der Waals in­ power plants as well as sweetening of natural gas, as they are heavily
teractions using highly brominated large anions such as 1-ethyl-3-me­ diluted with N2 and also contain other impurities. Hence, the relative
thylimidazolium hexabromophosphate (i.e., [emim][PBr6]) that led to selectivity of CO2 especially with respect to N2 is most relevant for carbon
enhanced solubility of CO2. Contrary to the claims of solvent-solute in­ capture from flue gas streams. Fig. 28 represents relative solubilities of
teractions governing solubility, others have suggested that entropic ef­ various industrially important gasses in the ionic liquid [hmpy][Tf2N] (1-
fects dominate CO2 solubility in ionic liquids [220]. This conclusion is n-hexyl-3-methylpyridinium bis(trifluoromethanesulfonyl)imide), and
supported by the observed increase in CO2 solubility with increasing hence the selectivity of CO2 with respect to other gaseous species. The
molecular weight, free volume, and molar volume of the ionic liquid solubility of gasses in [hmpy][Tf2N] at room temperature increases in the
following order, N2<O2<CH4<C2H6<CO2<SO2 [216]. A similar trend is
observed also with other ILs. In general CO2 solubility is significantly
higher than for N2 (and CH4), which makes it attractive for these ionic
liquids to serve as physical solvents to capture CO2 from power plant flue
streams, and also for the sweetening of natural gas.

Fig. 27. CO2 solubility in various imidazolium-based ionic liquids at 25 ◦ C,


where filled symbols denote ionic liquids with hydrogen attached to the 2-car­
bon position while the open symbols represent ionic liquids with methyl group
substituted for the 2-carbon position.
Filled square: [emim][Tf2N] or 1-ethyl-3-methylimidazolium bis(tri­
fluoromethylsulfonyl)imide; empty square: [emmim][Tf2N] or 1-ethyl-2,3-
dimethylimidazolium bis(trifluoromethylsulfonyl)imide; filled circle: [bmim] Fig. 28. Comparison of solubilities of various industrially important gasses in
[PF6] or 1-n‑butyl‑3-methylimidazolium hexafluorophosphate; empty circle: the ionic liquid [hmpy][Tf2N] (or, 1-n-hexyl-3-methylpyridinium bis(tri­
[bmmim][PF6] or 1-n‑butyl‑2,3-dimethylimidazolium hexafluorophosphate; fluoromethanesulfonyl)imide) at 298 K. High solubility of CO2 provides good
filled triangle: [bmim][BF4] or 1-n‑butyl‑3-methylimidazolium tetra­ selectivity for separation from typical flue gas compositions. The ordering of
fluoroborate; empty triangle: [bmmim][BF4] or 1-n‑butyl‑2,3-dimethylimida­ solubilities follows the trend N2<O2<CH4<C2H6<CO2<SO2 (in the inset).
zolium tetrafluoroborate. Reprinted from [216] with permission from the American Chemical Society.
Reprinted from [217] with permission from the American Chemical Society.

25
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

of the alkyl group.


Needless to say, the prospects of ionic liquids for CO2 capture ap­
plications require in the least competitive, if not superior, attributes
when compared to commercial solvents currently available for this
purpose (see Table 10). Although ionic liquids generally possess superior
properties including extremely low vapor pressure and high selectivity
for CO2/N2 mixtures, more work is required to develop ILs with more
competitive properties than other currently available solvents. ILs are
also quite expensive and sell for ~$1000/kg [216]. Mass production is
expected to bring the price down to about $40/kg, but this is still 10 to
20 times more expensive than the commercially available solvents.

5.2. CO2 capture by solid sorbents

Compatibility between the high exit temperatures of flue gasses and


operational parameters of the CO2 capture process is highly desirable for
effective thermal management and increased process efficiency. Capture
by absorption in liquid solvents requires cooling of the flue gas stream
Fig. 29. Relation between viscosity and CO2 diffusivity in various solvents,
down to temperatures that are compatible with the chemical stability
which seem to be proportional to (viscosity)− 0.43. Reprinted from [216] with
permission from the American Chemical Society.
and vapor pressure of the solvent used. By contrast, solid sorbents are
usually tolerant to and compatible with high temperatures of flue gasses
emitted from industrial processes and power plants, chemically ore
Another important consideration from a practical point of view is the
stable, and generally less expensive than liquid solvents. These attri­
generally high viscosity of ionic liquids, which can range between 66
butes make solid sorbents attractive candidates for CO2 separation and
and 1100 cP at room temperature [186]. These viscosity values are
capture.
significantly higher than most of the common solvents, which of course
As depicted in Fig. 30, a wide variety of solid sorbents including
adversely impacts interfacial contact between the gaseous and liquid
zeolites, carbon- and calcium-based sorbents, organic-inorganic com­
phases during absorption. In fact many ionic liquids can form highly
posites, and metal-organic framework structures have been investigated
viscous gel-like structures. Naturally, high viscosity is not a desirable
for their CO2 sorptive properties and are reviewed in recent articles [39,
property because it often limits mass transport of diffusing species, and
78,184,190,224,225]. In general, CO2 capture by solid sorbents pro­
poses practical challenges for the separation process and construction
ceeds in several ways. CO2 may be selectively adsorbed on the solid
materials. Fig. 29 depicts the relationship between the self-diffusion
surface (i.e., adsorption), it may chemically react with the solid to form a
coefficient of CO2 in various common solvents and ionic liquids with
stable carbonate (i.e., carbonization or mineralization – for storage), it
their viscosities, and indicates significant reduction in diffusion coeffi­
may diffuse into the pores of the solid (i.e., membrane separation), or its
cient with increasing viscosity. Fortunately, the properties of ionic liq­
diffusion in the solid is controlled by size selection rules (i.e., molecular
uids can be tuned by the proper choice of anions and cations. Generally,
sieving). In selecting the proper solid sorbent, consideration for desir­
IL viscosity increases with the size of the anion, and also with the length
able attributes includes high CO2 sorption capacity, high selectivity for

Fig. 30. Cartoon depicting examples of the rich variety of solid sorbent families of materials for CO2 capture. Reprinted from [224] with permission from John Wiley
and Sons.

26
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

CO2, fast adsorption/desorption kinetics, ease of regeneration with low adsorption of CO2. Summary of select physicochemical properties of
energy requirement, operational temperature and pressures compatible various gasses typically found in flue gas streams are compiled and
with flue gas conditions, high recycling stability, and chemical and presented in Table 11. The nature and extent of electrostatic interactions
thermal stability against and flue gas contaminants such as SOx, H2O, of these gasses, and in particular N2 and CO2, with the sorbent surface
HCl, NOx, and Hg. Unfortunately, there is no single sorbent that em­ sites not only affect the affinity for CO2 but also impacts the energy
bodies majority of these attributes. Inevitably, there are always trade requirements for the capture and regeneration processes. If the inter­
offs. On the other hand, materials costs for most solid sorbents are action is too strong, desorption becomes an energy intensive process. On
usually low. Nevertheless, it is estimated that the cost solid sorbent the other hand, while a weak interaction helps reduce energy demand
should not exceed $10/kg in order to compete favorably with liquid for regeneration, it also reduces CO2 selectivity.
solvents [190]. There are several key requirements for sorbent selection. The
adsorption capacity of the solid sorbent is a critical parameter and is
5.2.1. Capture by physical adsorption determined primarily by two factors, namely the specific surface area
The CO2 capture capacity by absorption in liquid solvents is gener­ and the equilibrium adsorption isotherm, both of which in turn dictate
ally limited by the liquid/gas interfacial contact area and rate of mass the required amount of sorbent and thus the size of the adsorption tower.
transfer across this interface. To achieve significant capture capacities, The higher the specific surface area, the higher the active surface site
this constraint necessitates the use of large area absorption towers and population available for adsorption. Moreover fast adsorption/desorp­
scrubbers. tion kinetics is desired for high capture throughput and fast cycling
Use of solid sorbents can overcome this limitation as solids with large between adsorption and desorption steps. High selectivity for CO2
surface areas can be fabricated to maximize the gas/solid contact area adsorption versus nitrogen and other flue gas components is highly
for adsorption. Moreover, mass transport rates (or, diffusion co­ desirable to obtain a concentrated CO2 product stream without the need
efficients) in liquids are several orders of magnitude slower than in for further separation. It is also desirable for the solid sorbent to have a
gasses. Typically, diffusion coefficients for dissolved or ionized species low heat of adsorption, as this determines the heat requirement for the
in liquids are of the order of 10− 9 m2/s, while in gasses the diffusion sorbent regeneration step, which may be substantial. For example, the
coefficients are generally ~10− 5 m2/s. This renders gas/solid interfacial heat for physisorption of gasses on solids usually ranges between − 25 to
rate processes to occur inherently faster and allows ready replenishment − 50 kJ/mol, while for chemisorption the range is between − 60 to
of adsorption sites by CO2. − 90 kJ/mol [190]. Between regeneration and adsorption cycles, which
However, for solid sorbents to compete with amine-based CO2 cap­ are typically carried out under different temperature and pressure re­
ture, they need to offer capacities higher than the average MEA scrub­ gimes, the sorbent is also required to exhibit good chemical, thermal and
bing capacity of 3 to 4 molCO2/kg of sorbent [190]. Despite the mechanical stability for prolonged use. During the
significantly faster diffusion rates of CO2 in the micropores of the solid adsorption-regeneration cycles, the adsorbed CO2 can be removed from
sorbent, CO2 capture by adsorption is impacted by interference effects the solid sorbent mainly in two ways, either by cycling the pressure, as in
particularly of water vapor and also by other gasses present in the flue pressure swing adsorption (PSA) and vacuum swing adsorption (VSA),
gas that compete for the same adsorption sites. CO2 adsorption involves or by cycling the temperature as in temperature swing adsorption (TSA).
interaction with a free active site, *, on the surface of the solid sorbent, These processes usually require two towers operating in tandem,
and regeneration involves the desorption of CO2 from that active site, each tower switching between adsorption and regeneration cycles out of
freeing * for the next adsorption cycle according to the reactions phase in time with the other tower. During the sorbent regeneration
step, the captured CO2 is recovered. The regenerated sorbent is then
CO2(g) + * = (CO2)(ads)*(adsorption step) (12) exposed to the flue gas for the next cycle of adsorption. Hence, the flue
(CO2)(ads)* = CO2(g) + *(desorption step) (13) gas flow is switched between these two towers. In PSA, CO2 from the flue
gas is preferentially adsorbed on the solid sorbent at high pressures
Selective adsorption of CO2 (versus N2) on the sorbent material is during the adsorption step, and released from the sorbent during the
governed by the strength of van der Walls interaction between the polar regeneration step at low pressure, or under vacuum (VSA) [225,227].
surface sites and the large quadrupole moment of the dipolar molecule
CO2 (4.30 D.A, or − 14.27×10− 40C.m2) compared to N2 (1.54 D.A, or
− 4.65×10− 40C.m2) [226]. Although neither N2 nor CO2 possess dipole Table 12
moments, the difference in their polarizabilities also favors preferential Comparative CO2 adsorption capacities of commonly used solid sorbents. AC:
activated carbon, SWCNT: single wall carbon nanotube, MWCNT: multi-wall
Table 11 carbon nanotube, MCM: anionic-surfactant-templated mesoporous silica, SBA:
Physicochemical properties of gaseous species commonly present in flue gas Santa Barbara Amorphous-type material, HMS: hexagonal mesoporous silica.
streams. Data are compiled from [226], and also from [89] and [183]. Modified from [184].

Species Kinetic Polarizability Dipole Quadrupole Sorbent Surface Adsorption Feed Gas Adsorption
Diameter (10–25 cm-3) Moment Moment Area Capacity Composition Temperature
(Å) (D) (D.Å) (m2/g) (mmol/g) (◦ C)

H2 2.89 8.04 0 0.66 Activated 1762 1.66 – 25


N2 3.64 17.4 0 1.54 carbon
O2 3.46 15.8 0 0.39 Meso- 798 1.50 100% CO2 25
CO 3.76 19.5 0.112 2.50 carbon
NO 3.49 17.0 0.159 SWCNT 1587 4.02 100% CO2 35
H2O 2.65 14.5 1.855 2.30 MWCNT 407 1.73 50% CO2
H2S 3.60 37.8 0.970 Graphene 1550 7.95 100% CO2 –78
CO2 3.30 29.1 0 4.30 MCM-41 1229 0.14 15% CO2 in N2 75
CH4 3.80 24.48 0 0.02 SBA-15 950 0.11 15% CO2 in N2 75
HMS 561 0.22 100% CO2 25
Note: D = Debye = 3.33564×10–30C.m. Meso-Al2O3 271 0.84 100% CO2 25

27
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

PSA is a commercial technology for CO2 capture from power plants, with
efficiencies in excess of 85% [162]. In TSA, heating the sorbent tower
with hot air or steam regenerates the solid sorbent, whereby the
adsorbed CO2 is released from the sorbent surface and recovered for
capture. Although the regeneration time for TSA is typically longer than
for PSA, the purity of the captured CO2 can be higher (> 95%) while CO2
recovery from the flue gas can exceed 80% [228]. A techno-economic
analysis reported that the operating cost for TSA for direct capture
from air was estimated to be around $80 to 150/ton of CO2 [229].
A vast variety of porous solids have been explored towards CO2
capture purposes. Typical CO2 sorption capacities and operating con­
ditions for many of these solid sorbents [184] are summarized in
Table 12. It is clear from the table that in general, the adsorption ca­
pacities of most solid sorbents at ambient conditions are well below the
average CO2 capture capacities of 3 to 4 molCO2/kg of sorbent for liquid
solvents such as MEA (see Section 5.1.1. Chemical absorption - above).
Comparatively low CO2 selectivity and uptake capacity are major
drawbacks for most of the untreated solid sorbents. To further improve Fig. 31. CO2 adsorption isotherms on activated carbon measured at various
their adsorption properties, various pretreatment strategies including temperatures. Reprinted from [190].
surface modification by amines for imparting chemical functionality at
the solid surface or by impregnation of CO2-active liquid sorbents such
as amines into the pores of solid sorbents have been applied to pro­
spective sorbents. Similarly, solid-liquid and solid-polymer composite
sorbents are explored for improved CO2 capacity and selectivity. that water vapor preferentially adsorbs on oxygen containing surface
Although carbon-based sorbents has relatively higher capacities than sites or functional groups on carbons, and this adversely impacts CO2
silica-based sorbents, they demand high pressures and low temperatures sorption capacity [236,237]. By contrast, surface modification or
for increased capacity. Silica-based sorbents, on the other hand, offer decoration by amine solutions improves CO2 adsorption capacity by
surface modification more readily than carbon-based sorbents because providing surface sites that are of basic character. Indeed, the capacity of
of their high concentration of surface-OH groups that facilitate chemical coke-derived activated carbon impregnated with diethanolamine was
modifications to or functionalization of the silica surface to enhance CO2 reported to be 5.63 molCO2/kg of adsorbent [238].
adsorption. Indeed, surface modification or impregnation of porous Zeolites represent another important family of solid sorbents that are
solid sorbents with amine-based chemicals such as polyethylenimine used in many industrial applications. Zeolites may occur naturally or can
(PEI), diethenolamine, (DEA), and tetraethylenepentamine (TEPA) have be made synthetically. They have microporous structures based on a
resulted in improved CO2 capacities [184]. Also, CO2 uptake capacity of silicate framework with well-defined and uniform pores whose size may
5.39 molCO2/kg of as-synthesized mesoporous silica (MCM-41) sorbent vary typically from 0.5 nm to 1.2 nm [239]. Substitution by Al for Si sites
impregnated with TEPA has been reported [230]. The basis for such in the framework structure imparts a negative charge that is compen­
improvement is the enhanced interaction between the CO2, which is sated by an alkali or alkaline-earth cation. The cations can be
acidic, and the active amine site on the modified sorbent surface, which ion-exchanged to tune the pore size within the framework, which allows
is basic in character. Enhanced adsorption is facilitated by the covalent gas separation by molecular sieving. Affinity to CO2 adsorption is also
bond formation between the amine site and the adsorbed CO2. Similarly, contributed by its large quadrupole that gives rise to strong electrostatic
alkali-metal or alkaline-earth oxides can be employed for surface interaction with the cation sites. Nevertheless, as the gas separation
modification and improved capacity by providing active surface sites of mechanism is largely provided by size selection, detailed discussion of
basic character that enhances CO2 adsorption. Activated carbon is one of zeolites are provided below in Section 5.2.3.1.
the cheapest, oldest and most widely used solid adsorbents for industrial
applications, and is also considered for CO2 capture from power plants 5.2.2. Capture by chemical reaction
[231]. It has high thermal, mechanical and chemical stability, and ex­ Capture by chemical reaction exploits the base-acid interactions
hibits low sensitivity for water vapor, which otherwise competes with between the acidic character of CO2 and the basic character of alkali- or
CO2 adsorption with adverse impact on capacity. However, CO2 is only alkaline earth metal oxides or carbonates. Use of amines for surface
weakly adsorbed on carbonaceous sorbents, resulting in poor selectivity modification or impregnation of solid sorbents similarly serves the
and low capacity. These shortcomings are mitigated by using advanced purpose of enhancing the basic character of surface sites. Chemical
processing techniques that greatly improve surface area and pore capture of CO2 from flue gasses using alkali metal carbonates (M2CO3,
structure by nanostructuring of the carbon sorbent, or by doping or where M: Li, Na, K) [240] can be expressed by the carbonation reaction,
decoration to increase the alkalinity of surface sites to enhance CO2
M2CO3 + H2O + CO2 = 2 MHCO3 (14)
uptake capacity [232] It is well known that processing, nanostructuring,
functionalization and doping play important roles in designing The enthalpy for the carbonation reaction is − 135 kJ/mol for M = Na,
advanced functional materials with superior properties [233,234]. This and − 141 kJ/mol for M = K. CO2 uptake is accomplished in the presence
approach is reflected in the high adsorption capacities for single-walled of steam, where CO2 and steam react with the alkali metal carbonate
carbon nanotubes and graphene oxide [235] for CO2 adsorption in sorbent at 60 to 110 ◦ C and form an alkali metal bicarbonate as indicated
Table 12, where the latter exhibits a capacity of 7.95 molCO2/kg of in reaction (14). The corresponding reaction for the regeneration (or,
sorbent. The table also points to the strong influence of temperature on decarbonation) of the solid sorbent can be expressed by,
capacity. As the heats of adsorption for most gasses on solid are
exothermic, it is to be expected that lowering the adsorption tempera­ 2 MHCO3 = M2CO3 + H2O + CO2 (15)
ture improve their CO2 uptake capacity. Indeed, adsorption isotherms of The regeneration step involves heating the bicarbonate at 100 to 200 ◦ C
CO2 on activated carbon illustrates this trend clearly in Fig. 31, where during which the CO2 is released and captured. The theoretical CO2
higher CO2 capacities are obtained at lower temperatures. Investigations capacities for Na2CO3 and K2CO3 are 9.43 molCO2/kg and 7.23 molCO2/
on various forms of nanostructured and mesoporous carbons indicated

28
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

kg of sorbent, respectively. These values are more than 2 to 3 times in Fig. 32, where K2CO3 exhibits the highest capacity for CO2 capture
higher than for sorbents that capture by physisorption. However in re­ among the other alkali metal carbonates.
ality, only a portion of the sorbent is actively utilized for CO2 uptake. Interestingly, lithium silicate exhibits higher CO2 capacity than its
Indeed, most studies reported capacities ranging roughly between 1 and zirconate counterpart, and up to 10 times the capacity with respect to
3 molCO2/kg of either Na2CO3 or K2CO3 supported on various solid other oxides, while offering rapid adsorption kinetics, stability, and
supports including activated carbon, porous Al2O3, silica gel, TiO2, and operation in a wide range of elevated temperatures and flue gas CO2
ZrO2 [190]. The results demonstrate the role of bulk diffusion limita­ concentrations [245]. The overall chemical reaction for CO2 capture is
tions of the carbonation reaction, and despite the dispersion of the active
sorbent on solid supports, that only a fraction of the sorbent mass can Li4SiO4(s) + CO2(g) = Li2SiO3(s) + Li2CO3(s) (17)
actively be utilized for CO2 capture. This limitation is one of the draw­ The lithium silicate reacts with CO2 below 720 C to form Li2CO3,

backs of chemical capture by solid sorbents. and then releases the CO2 when heated above 720 ◦ C. This narrow
temperature window offers great thermal advantage, and is arguably
5.2.2.1. Capture on oxides and minerals. Capturing of CO2 by reacting one of the highest operating regimes for solid capture including mo­
with alkaline earth oxides or carbonates such as calcite (CaCO3) or lecular sieves, membranes, and metal-organic frameworks structures.
dolomite (CaCO3.MgCO3) have long been used for sweetening of natural Naturally, these attributes potentially make Li4SiO4 a viable option of
gas and other flue streams for the removal of CO2 and sulfurous species post-combustion CO2 capture by carbonization. However, the overall
H2S and SO2 [241]. This approach potentially lends itself to calcium rate of CO2 uptake, which exhibits first order kinetics with respect to
looping process (CLP) at high temperatures compatible with power plant CO2 partial pressure, seems to be controlled by a highly activated
flue gas streams [242]. Although these sorbents are cheap and abundant, (~120 kJ/mol) and sluggish surface reaction [246].
they suffer from loss of more than half of their capture capacity during
repeated carbonation-decarbonation cycles. Moreover, the rate of CO2 5.2.3. Capture by molecular sieving
uptake and CO2 capacity of the solid is severely hindered by mass Separation by molecular sieving is achieved by framework solids that
transport effects inside the bulk. This limits the usable capacity of these have crystal structures with sufficiently large open channels or pores of
cheap and abundant materials as an effective CO2 sorbents. Doping the order of nanometers in size that are compatible with molecular di­
strategies to create lattice defects inside the crystal help enhance mass ameters of gasses. Depending upon the type and composition of the
transport rates in the solid sorbent. Many of the issues related to the framework solid structure, these channels selectively allow size-
fundamental aspects of the capacity decay in the calcium-based sorbents compatible gas molecules to pass through but not others. Transport
are reviewed recently, where hydration during or after the calcination (i. mechanism through these solids is generally via Knudsen diffusion.
e., decarbonation) step is proposed to reduce the loss in capacity [243]. In many solids of this category, not only the size but also the shape of
Other oxides were also investigated for post-combustion CO2 capture the channel is important and permits shape selective catalysis [247]. An
in a similar looping process. As an example, zirconates (Li2ZrO3) and important industrial example of the latter is the commercial
silicates (Li4SiO4) of lithium such are reported to show promise for high methanol-to-gasoline (MTG) conversion process developed in the 1970′ s
CO2 sorption capacity at elevated temperatures, where the capture re­ by Mobil (now, ExxonMobil) using shape selective zeolite (ZSM-5) cat­
action can be expressed as, alysts to convert methanol to gasoline.
Li2ZrO3(s) + CO2(g) = Li2CO3(s) + ZrO2(s) (16)
5.2.3.1. Capture on zeolites. Zeolites are a rich class of materials, mostly
Depending on the CO2 content of the flue gas, reaction (16) offers made of crystalline aluminosilicates, where the framework crystal
reversibility within the temperature range of 450 to 590 ◦ C [78]. This structure is built on the periodic arrangement of SiO4 and AlO4 tetra­
allows improved thermal management during switching between the hedra with well-define molecular size pores or channels, as illustrated in
capture and regeneration cycles simply by swinging the temperature in a Fig. 33. The framework structure is negatively charged due to the dif­
narrow window to reverse the reaction direction. Addition of other al­ ference between the oxidation states of Si+4 and Al+3 ions. This negative
kali metal oxides or carbonates such as K2CO3 also help improve the CO2 charge is compensated by the positive charge of exchangeable alkali-
uptake rate and capacity by forming binary or ternary eutectic compo­ metal or alkaline-earth cations located in the pore spaces.
sitions on the surface of the zirconate particles, whereby accelerating the Gas separation by zeolites depends on several factors but is governed
absorption of CO2 in this molten carbonate film [244]. This is depicted largely by the structure and composition of the zeolite, the nature and
size of structural cations located in the pores, and the size, shape and
polarity of the individual gas molecules. The alkali cation sites that are
basic in character are primarily responsible for facilitating adsorption of

Fig. 32. Comparison of CO2 uptake by alkali metal carbonates as a function of


temperature, for a feed gas composition of 13.8% CO2, 10% H2O, and balance
He. Reprinted from [190] with permission from the American Chemi­
cal Society.

Fig. 33. Framework structure of a zeolite with well-define pores [248].

29
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

the acidic CO2 molecule. Because of the larger quadrupole moment of Table 13
CO2 (− 14.27×10− 40C.m2) compared to N2 (− 4.65×10− 40C.m2) [226], CO2 and N2 sorption capacities and selectivities of chabazite (CHA), Linde Type
there is also an electrostatic contribution due to stronger interaction of A (A), and ZK-5 families of zeolites. Values for CO2 uptake is for 15 kPa pressure,
CO2 with the structural cations in the pores of the zeolite framework that while it is 85 kPa for N2, both measured at 273–298 K. Also, separation factor (s)
favors adsorption. Hence relative polarizabilities of CO2 and N2 play a is defined as (CO2 uptake/N2 uptake)/(15 kPa/85 kPa) [253].
role in effective separation, i.e., its selectivity. Adsorbent CO2 uptake N2 uptake Selectivity(s)
There are more than 170 unique zeolite topologies whose structures (molCO2/kg) (molN2/kg)

have been approved by the Structure Commission of the International Li-CHA 4.4 0.53 47
Zeolite Association (IZA-SC) [250]. Because of the vastness of zeolite Na-CHA 4.2 1.3 18
K-CHA 4.0 0.85 27
compositions and structures, only the salient features of CO2 adsorption
Mg-CHA 3.4 0.65 30
on these materials are discussed below. For more details the reader is Na-A 3.2 0.30 60
referred to the vast literature and review papers on zeolite structures as Na-A (17% K+) 2.3 0.02 660
well as the IZA web site [225,250–255]. Mg-A 2.4 0.25 54
Some of the common features of zeolites are: (a) they have Ca-A 4.0 0.50 45
Li-ZK-5 3.9 0.23 96
exchangeable (usually, alkali metal) cations residing inside the pores
Na-ZK-5 3.4 0.27 71
that can be tuned for enhanced catalysis or CO2 capture, (b) these cat­ K-ZK-5 3.0 0.23 74
ions, when exchanged with hydrogen (i.e., H+) provides strongly acidic Mg-ZK-5 1.9 0.15 72
sites for catalysis but not for CO2 capture, (c) they have pores or chan­
nels that are generally less that 1 nm, which is compatible with the
effective kinetic diameters respectively, of 0.33 nm and 0.36 nm for the
CO2 and N2 molecules inside the zeolite pores [254], and (d) these pores
may have one or more discreet sizes depending on the type of alkali
cation present. The last two features are critical for size selection nature and size of the cations are important considerations as they
capability because they may allow separation by molecular sieving. impact the pore size and acid-base properties for the CO2 adsorption site.
Monovalent cations are typically large, so they tend to block a large For example, CO2 capacities for ion exchanged NaX and NaY zeolites
section of the pore cross-sectional area and reduce the pore size to below indicated higher capacity with decreasing cationic radius in the order
0.4 nm, which limits passage to only small molecules, while alkaline Cs+<Rb+<K+<Na+<Li+, such that at ambient conditions of tempera­
earth or other divalent cations occupy every other cation position ture and pressure, CO2 capacities of Li+-exchanged X and Y achieved 5.6
leaving some pores available for gas transport of larger molecules. Be­ and 5.2 mmolCO2/g of zeolite while the corresponding capacities for
sides capacity, the rate of CO2 uptake is operationally a critical Cs+-exchanged X and Y zeolites were 3.1 and 2.6 mmolCO2/g of zeolite,
parameter, and this is dominated by diffusion through the micropores of respectively [255].
the zeolite crystal framework. Due to its larger kinetic diameter, diffu­ Comparative sorption capacities of CO2 and N2 as well as the
sion of N2 to access adsorption sites inside the smaller micropores is resulting separation factors for several zeolite families are summarized
comparatively restricted and hinders N2 uptake in the zeolite structure. in Table 13. Although most compositions exhibit high selectivity, the
Adsorption isotherms provide a measure of the CO2 adsorption ca­ uptake values indicate modest sorption capacities in the range of 2 to 4
pacities of zeolites as illustrated in Fig. 34 for a variety of commercially molCO2/kg of zeolite, barely competitive with amine-based sorbents
available zeolite compositions as a function CO2 pressure. As expected, discussed above. Nevertheless, many zeolite families both synthetic and
adsorption capacity increases with increasing pressure. Composition naturally occurring have been developed commercially for multitude of
also plays an important role even within the same family. Indeed, the industrial applications. For example, zeolite 13X was successfully
employed for both VSA [225] and PSA applications, where in the latter
13X was shown to recover 99% pure CO2 from 53% of the CO2 from low
CO2-content (16% CO2 balance N2) flue gas and 70% from high-­
CO2-content (26% CO2 balance N2) gas mixtures [256]. Depending on
type, the synthesis of zeolites can be process-intensive and costly, but the
raw materials are usually inexpensive. However, zeolites are usually
hydrophilic and show preference for H2O over CO2. This makes
capturing CO2 from moist environments or flue gasses quite challenging.
The primary advantage of zeolites is their ability to tune the composition
and tailor towards specific properties. The strong interaction between
the adsorbed CO2 and the cation site that is responsible for the high CO2
uptake capacity, at the same time, hinders the release of CO2 during
cycling and regeneration. By choosing the right cation substitution, one
can reduce electrostatic interactions between the adsorbed CO2 and the
cation and mitigate this problem, but this comes at the expense of lower
CO2 uptake capacities.

5.2.3.2. Capture on metal organic frameworks (MOF). Metal-organic


framework (MOF) structures [257,258] have gained interest as a
promising class of new materials with potential applications in catalysis
[259,260], hydrogen and methane storage, CO2 capture [183,261], and
gas separation [262,263]. However, they are yet to be demonstrated for
Fig. 34. Adsorption isotherms for CO2 on various commercially available ze­ CO2 capture at pilot scale [264]. MOFs are made of two types of building
olites, which allow a measure of capture capacity Reprinted from [249] with units, namely, a transition metal ion or a cluster of ions and organic
permission from Elsevier. linkers that connect them. Flexibility in changing these building units
allows modification and tuning of the structural and functional

30
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 35. Schematic illustration of a representative MOF structure, showing the


metal-organic framework structure with building units surrounding a large void
space at the center indicated by the large yellow sphere. Reprinted from [257]
with permission from the American Chemical Society.

Fig. 36. Isotherms for CO2 adsorption on various MOF compositions at ambient
temperature. Reprinted from [266] with permission from the Royal Society
of Chemistry.
properties of MOFs. The building units form three-dimensional organ­
ic-inorganic hybrid networks via multiple links between transition metal
ions such as Cu and Zn and organic ligands including a variety of car­
boxylates, azolates and phosphonates. The structure and shape of the highest reported CO2 uptake capacity of about 55 molCO2/kg of
metal-organic assembly is governed largely by the nature of metal ions MOF-210 at room temperature and 5 MPa pressure [269].
and organic ligands, and how they are linked together in the framework A recent addition to the class of framework structure materials is the
structure. The length of the organic ligand affects the accessible surface zeolite imidazole framework (ZIF) materials where the metal atoms such
area and hence the porosity, which increases the accessible adsorption as Zn are linked through N atoms imidoazolate (C3N2H3− : Im) or func­
sites and improves adsorption capacity. However, long ligands result in tionalized Im links. The topology of ZIFs is largely governed by the
structurally more fragile MOFs. functionality of the imidazolate, which allows topology modifications
The framework structure allows MOFs to form a wide variety (more achieved by specific functionalization of the Im group. This is illustrated
than 20,000) of topologically diverse structures, functional groups, pore in Fig. 37, where the pore size and surface area display almost a linear
sizes and porosities. The cartoon of a MOF structure is illustrated in relationship. Such controllable flexibility in varying the pore size and
Fig. 35, where the large yellow sphere in the center represents the void surface area, as well as their thermal stability up to 300 ◦ C [267] makes
space, or the pore. Accordingly, MOFs are highly porous materials with ZIFs an attractive class of materials for gas separation, catalysis and
pores up to 9.8 nm and densities down to 0.13 g/cm3 [2560]. Indeed other application.
very large specific surface areas up to 10,000 m2/g of MOF can be ob­ Despite such high CO2 capacities and tunable properties such as pore
tained [265]. Such extraordinarily high surface areas and chemically size and surface area, it is too early to assume that MOFs (or, ZIFs) may
adjustable pore sizes allow tunable internal surface functionalities and lead to promising solutions for post-combustion CO2 capture. In fact,
properties that provide potential for a wide range of applications thermal and mechanical stability at temperatures significantly higher
including CO2 capture. The heat of adsorption of CO2 for MOFs is than room temperature as well as thermal compatibility with flue gas
generally low and comparable to those for zeolites. Values in the range streams are general concerns. As with most solids, their CO2 adsorption
of − 30 to − 45 kJ/mol for MIL-53 (Al, Cr) from 0.1 to 0.4 MPa pressure, capacities rapidly decrease with increasing temperature and decreasing
and − 30 kJ/mol for USO-1-A1 from 0 to 0.04 MPa pressure were re­ CO2 pressure. Hysteresis in capacity between adsorption and desorption
ported [224]. The metal-organic framework composition MOF-5 is one cycles is also not fully understood. The impact of moisture and other coal
of the very first synthesized and widely studied cubic framework group contaminants on the structural and chemical stability as well as the
of MOFs with the formula [Zn4O(BDC)3]•(DMF)8(C6H5Cl), where adsorption capacity requires further investigation and improvement for
BDC=1,4-benzenedicarboxylate, and DMF=N,N’-dimethylformamide. post-combustion CO2 capture [268]. There is of course room for progress
Many other MOF structures and compositions have been synthesized in improving the adsorption capacity especially at low CO2 pressures by
and investigated for CO2 adsorption since then [183,258,262,266–268]. synthesizing MOFs with proper pores sizes and surface area, as well as
The adsorption capacities of several MOF types measured mostly by by strengthening the interactive forces between the CO2 molecules and
thermogravimetric analysis (TGA) are presented in Fig. 36. The iso­ MOFs by introducing unsaturated metal sites in the structure. Further­
therms conform to a Langmuir-type shape, where the best performance more, devising cost-effective synthesis routes for MOFs can improve
was obtained with MOF-177 above 1.5 MPa pressure. At lower pressures their economic competitiveness. These concerns and knowledge gaps
of 0.1 MPa, MOF-74, MOF-505, and Cu3(BTC)2 show the highest CO2 require full understanding and resolution before MOFs and ZIFs can be
capacities among the nine MOFs presented in the figure. However, employed for post-combustion CO2 capture.
MOF-210 (not in the figure) that is reported to have a bulk density of
0.25 g/cm3 and a specific surface area of 6400 m2/g provided one of the

31
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 37. Linear relationship between pore diameter (dp) and surface area of GME ZIFs governed by the kno cage shown for each composition. C: black, N: green, O:
red, Cl: pink, Br: brown, Zn: blue tetrahedral (H atoms are not shown for clarity). Reprinted from [267] with permission from the American Chemical Society.

5.3. CO2 separation by membranes no solubility for CO2 and hence not useful for CO2 separation. Neither
are MIEC oxide membranes that are selective for oxide ion transport
Gas separation membranes have been employed in the industry for only (see Section 4.1.3), and not suitable for separation of CO2. Table 14
nearly three decades, mostly for air separation or oxygen enrichment provides a summary of major types, properties and operating principles
applications aimed for combustion processes. Recently, oxygen enrich­ of different membrane families that are relevant to separation of CO2 for
ment or purification by air separation membranes have also gained in­ capture.
terest for thermoelectric power generation by integrated gasification The two most important parameters for both polymeric and micro­
combined cycle (IGCC) and oxy-combustion processes [111,138,270]. porous membranes are selectivity and permeability, the product of
Despite the technological gains in O2 separation, however, there are which is defined as the separation factor (also called, separation power).
currently no commercial membranes available for CO2 separation and
Sepration Factor = Selective × Permeability (18)
capture from power plant flue gas streams, but there is ample interest
[89]. Naturally, high separation power is a desirable quality of a pro­
Major advancements in the development and commercialization of spective membrane. However, membrane selectivity and permeability
gas separation membranes in the 1980′ s were achieved by polymer- unfortunately have an inverse relationship, whereby high permeability
based asymmetric membranes used primarily for oxygen enrichment results in low selectivity, and vice versa. This poses a fundamental
via air separation [271]. In addition, mixed matrix or polymer-ceramic barrier to achieve high selectivity and high permeability at the same
composite membranes have been explored for enhanced selectivity and time with the same membrane type or composition. Tuning and modi­
permeability [125]. Similarly, ceramic membranes with microporous or fication of materials properties by doping, pore size and porosity con­
mesoporous microstructures have been developed. Microporous mem­ trol, and process engineering, as well as surface treatment and
branes may be employed as such or impregnated with a suitable functionalization may help achieve more favorable membrane proper­
CO2-active secondary phase such as an amine solution. Mesoporous ties. In both polymeric and microporous membranes, the driving force
membranes such as zeolites and metal-organic framework structures for transport is the pressure difference, i.e., the chemical potential dif­
have well defined crystallographic channels and voids that allow mo­ ference, of the gaseous species across the membrane. Naturally, larger
lecular sieving. On the other hand, dense inorganic membranes made the pressure differential and thinner the membrane, steeper is the force
from Pd-alloys such as Cu-Pd and Ag-Pd have absolute selectivity for H2 gradient across the membrane wall and higher is the mass flux, or
through ionic transport through the Pd crystal lattice and have been permeability. So it is desirable to fabricate the membranes as thin as
widely used in the industry for separation and purification of H2 espe­ possible, but also with adequate mechanical integrity to withstand and
cially from gas mixtures of water gas shift reaction products, but possess maintain the pressure differential across the wall.

Table 14
Types and operating principles of membranes relevant for CO2 capture from flue streams. Revised from [89].
Membrane Types Material Types Operating Principle CO2 Selectivity Gas Permeability Thermal tolerance

Polymer Glassy Gaseous dissolution/molecular diffusion Moderate Low Low


Rubbery/ Low Moderate/
elastomeric High
Metal Pd-alloys Solution/ionic diffusion Absolute for H2, Low Moderate
none for CO2
Molten Carbonate Carbonate eutectics Ionic transport Absolute for CO2 Moderate High
Porous Inorganic Silica, Carbons Sorption/diffusion Moderate Moderate High
Impregnated ceramics High Moderate Moderate
Molecular Sieving Zeolites Sorption/diffusion High Moderate Moderate
MOFs, ZIFs High Moderate Low

32
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

5.3.1. CO2 separation by polymeric membranes necessary.


The dominant fundamental mechanism for gas separation in poly­ Thermal and mechanical stability of the polymeric membrane are
meric membranes is by preferential solution-diffusion of gaseous spe­ also critical considerations. High exit temperature of hot flue gasses
cies, whereby certain gas molecules absorb or adsorb within void spaces from the power plant is not compatible with the operating temperature
in the polymer matrix and diffuse through the open channels among the regime of polymeric membranes, and hence the flue stream must be
polymer chains, while other components of the gas mixture pass through cooled preferably down to less than 100 ◦ C in order to avoid thermal
membrane relatively less affected. This is due to interaction of individ­ damage to the polymeric structure of the membrane. At the same time,
ual gaseous species with the polymer matrices with different strengths the flue gas needs to be pressurized to generate a pressure gradient
and rates for absorption, adsorption and diffusion. Accordingly, absolute across the membrane that drives the separation and collection of CO2 on
selectivity for a certain gas in a mixture is almost impossible to achieve the permeate side. Nevertheless, progress has been made in both fronts.
with polymeric membranes. There is always a trade-off between selec­ A polymeric membrane made of poly(vinyl alcohol) (PVA)-poly(acrylic
tivity and flux (or, permeability). Typically high permeability polymers acid) (PAA) copolymer and blended with the selective CO2 carrier 2,3-
such as rubbery or elastomeric open structure polymers offer high diaminopropionic acid hydrochloride (DAPA-HCl) not only improved
permeability but poor selectivity, while glassy or semi-crystalline CO2/H2 selectivity but also the CO2/N2 selectivity (= 700) and CO2
polymers offer high selectivity but low permeability. For this reason, permeability (2 × 10− 4 mol/m2.kPa.s) at a relatively high temperature
the gas mixture needs to be recycled many times through the membrane (160 ◦ C) and pressure (600 kPa) [272]. Nevertheless, these thermal and
in order to achieve the required level of enrichment. In other words, mechanical operations and steps add cost and energy penalty to the
polymeric membranes cannot accomplish absolute separation of CO2. overall capture process. The latter needs to be paid by the produced
The difficulty of selective separation of CO2 from N2-rich mixtures is energy and naturally lowers the overall plant efficiency as well as for
highlighted by comparison of their closely related physicochemical CO2 capture. Indeed, a recent study compared post-combustion CO2
properties that are presented in Table 11 above. Most importantly, the capture by polymeric membranes versus amine absorption and
kinetic diameter of CO2 (0.33 nm) is quite close to N2 (0.364 nm), which concluded that from an energetic point of view, membrane separation is
makes it difficult to use size exclusion effectively. So further develop­ not competitive with amine absorption, and estimated that for 10% CO2
ment of CO2-selective polymeric membranes is likely to follow two in the flue gas, the energy requirement for polymeric membrane may be
strategic approaches, namely, increase the CO2 solubility and enhance around 8 MJ/kgCO2, while it is about 4 MJ/kgCO2 for amine absorption
CO2 diffusion rate through the polymer matrix. Careful design and [273].
fabrication of polymers with appropriate void spaces or channels having
a certain size, side groups, structure, and chain conformation may 5.3.2. CO2 separation by microporous inorganic membranes
improve both selectivity and permeability. Chain dynamics alter the In general, inorganic membranes can be viewed in two broad cate­
void spaces in the polymer matrix and allow the diffusion of the gaseous gories in terms of their porosities, namely, porous membranes, and
species. To achieve improved selectivity and permeability in polymeric impervious (or, dense) membranes [274]. However, only porous inor­
membranes, mixed-matrix membranes made by grafting or blending ganic membranes offer practical relevance to CO2 separation. Major
different polymers having complimentary properties, or by blending advances in inorganic membranes for CO2 separation and capture have
inorganic adsorbents into polymer matrices have been explored [125]. been discussed and presented in recent review articles [89,275]. These
Partly due to its smaller kinetic diameter of 0.265 nm (see Table 11), membranes may be made of various carbon nanostructures such as
H2O is preferentially separated at the expense of CO2 through most carbon nanotubes or hollow fibers, or various oxides such as zeolites,
polymer membranes. This necessitates drying of the flue gas stream silica, alumina, and metal-organic frameworks. Electrochemical molten
before the membrane separation process. Selectivity and permeability membranes based on predominantly ionically conducting eutectics of
parameters for CO2 separation from CO2-N2 mixtures using various alkali-metal carbonates constitute another category discussed below.
single phase or blended polymeric membranes are presented in Table 15 Microporous inorganic membranes can be self-standing, or they may be
for comparison. deposited as thin layers or dispersions on inert or active substrates that
Aside from the mechanistic aspects of separation, CO2 capture by provide mechanical support and robustness. Support materials may be
polymeric membranes from power plant flue streams pose other chal­ made of porous ceramics such as silica, alumina, titania, zirconia, porous
lenges. The low concentration of CO2 in the flue gas stream requires glasses, or porous metals such as stainless steel. As an example, micro­
large membrane surface area to process huge volumes of flue gas. The structure of a zeolite membrane supported on a porous alumina tubular
membrane material must have sufficient chemical stability against the substrate is presented in Fig. 38. The inherent chemical and thermal
wide range of contaminants present in the flue gas streams emitted from stability of inorganic materials make them attractive for CO2 capture
coal-fired plants. Moreover, the flue gas requires drying prior to the from power plant flue streams. Their exceptionally wide operating
membrane separation operation as water vapor in the flue gas competes temperature regime from room temperature up to 1000 ◦ C or above is a
favorably with CO2 and diffuses faster through the membrane. Hence, a major advantage for thermal management of the flue gas feed directly
gas clean up operation prior to the membrane separation process is from the power plant and results in high efficiency for CO2 capture.
Despite their chemical stability and resilience against flue gasses con­
taining coal contaminants, presence of humidity in the feed gas
Table 15 adversely impacts the adsorption and selective transport properties of
CO2 separation characteristics of various polymeric membranes [78].
most inorganic membranes. Moreover, even though the membrane
Membrane Material Permeance Selectivity materials can withstand operation at elevated temperatures, the selec­
(m3/m2.Pa.s) (CO2/N2) tivity for CO2 diminishes at these temperatures. Hence, the optimum
Polyimide 735 43 operating temperature for most inorganic and composite membranes is
Polydimethylphenylene oxide 2750 19 below 200 ◦ C. Indeed, high permeability (10− 6 mol.m − 2.Pa− 1.s − 1 and
Polysulfone 450 31
high CO2/N2 selectivity (= 500) measured at room temperature for
Polyethersulfone 665 24.7
Poly(4-vinylpyridine)/Polyehterimide 52.5 20 zeolite Y-based microporous membranes reduced dramatically above
Polyacrylonitrile/Poly(ethylene glycol) 91 27.9 200 ◦ C, where selectivity enters the diffusion-controlled regime and the
Poly(amide-6-b-ethylene oxide) 608 61 relative difference between the diffusion coefficients of CO2 and N2 di­
minishes dramatically [276].
Gas transport mechanisms across microporous inorganic membranes
can generally be considered in four groups, namely, Knudsen diffusion

33
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 38. Porous support for a zeolite membrane. From left to right: alumina hollow fiber tube and electron microscope cross section showing the zeolite (MFI)
crystals filling in the micropores. Reprinted from [89] with permission from the American Chemical Society.

slows their development efforts.

5.3.2.1. Carbon-based membranes. Carbonaceous adsorbents such as


activated carbon have long been employed for many industrial separa­
tion processes due to their advantages including abundance, wide
availability, thermal stability, low cost, and low sensitivity to moisture.
On the other hand, despite the hydrophobic nature of most carbons,
moisture adsorbs competitively with CO2 on carbonaceous sorbents that
interact weakly with CO2 due to its relatively low heat of adsorption.
Comparatively, the heat of adsorption for CO2 on activated carbon is
− 30 kJ/mol, while it is − 36 kJ/mol for most zeolites [190]. Moreover,
the selectivity for preferential adsorption of CO2 on carbon-based ad­
sorbents and membranes decreases with increasing operating tempera­
ture, and this limits their practical prospects for CO2 capture from hot
flue streams.
Hence, much of the recent progress to improve the selective
adsorptive properties of carbon-based membranes has focused on
increasing surface area, controlling pore size distribution, nano­
structuring, and surface engineering including functionalization or
chemical activation to increase surface alkalinity using CO2-active ad­
ditives such as amines [190,279]. Several types of microstructures are
employed in carbon membranes, including single-wall and multi-wall
Fig. 39. Cartoon representing the role of the pore size governing the dominant carbon nanotubes, carbon molecular sieves, carbon hollow fibers, and
mechanisms for gas transport through inorganic microporous membranes. graphene [232,235,280,281].
Reprinted from [78] with permission from Elsevier. Many of the carbon microstructures for membranes are generally
made by pyrolysis of polymeric or organic precursors in the absence of
oxygen at elevated temperatures typically in the range 500 to 1000 ◦ C.
Generally, thermosetting type polymers are the preferred precursor
when pore size and the molecular diameter of the gaseous species are materials for this purpose. The resulting carbon microstructure, such as
comparable, surface diffusion where one gaseous species preferentially hollow fibers, can be supported on porous inert substrates including
has a higher surface diffusion coefficient, capillary condensation within alumina, zirconia, titania, or polymers for improved mechanical integ­
the pores of the membrane, and molecular sieving by size exclusion rity. The microstructures usually have narrow size distribution of in­
principle. Several of these transport mechanisms are schematically ternal bore diameters that can range from less than 1 nm to several
illustrated in Fig. 39. Diffusion within intersecting channels as in zeolite nanometers depending on the polymer precursor type and shape, as well
structures or between cages as in MOFs has also been considered as as the pyrolysis conditions. Such control of narrow size distribution
important transport mechanisms [89]. makes it possible to separate gasses by molecular sieving. Indeed, mo­
Gas separation by inorganic microporous membranes are dictated by lecular sieving was recently demonstrated using several stacked nano­
several factors including the heat of adsorption, which is indicative of sheets of graphene oxide that are supported on polyether block amide
how tightly or loosely the gas molecules bind to the membrane surface, (PEBA) copolymer substrate via hydrogen bonding, and gave superior
the sizes of gas molecules with respect to the pore size of the membrane, CO2 separation performance with CO2 permeability of 100 Barrer (1
porosity and tortuosity of the membrane, thickness of the active mem­ Barrer = 10− 12 m3(STP).m/(m2.s.mHg)) and selectivity of 91 from CO2/
brane layer, operating temperature, and specific surface area. Tuning N2 gas mixtures at room temperature and 0.3 MPa pressure [235].
the adsorption properties of the membrane by surface engineering [277] Similarly, it was reported that single-wall carbon nanotubes exhibit
and/or functionalization especially by surface amine groups [278] help twice the CO2 capacity of activated carbon [282].
improve the CO2 adsorption capacity and surface diffusion. Similarly, It should be noted that the chemical and thermal tolerance of carbon-
micro- and nanostructuring of the membrane by advanced processing based materials and the ability to synthesize or fabricate them in a
tools and techniques can allow tuning of the pore size and porosity controlled manner, make them attractive also for CO2 separation by
distribution in the membrane that can maximize the preferential pressure swing adsorption [256]. As adsorptive capacity for CO2 im­
transport of the desired gaseous species such as CO2. Nevertheless, proves with external pressure, carbon-based sorbents and membranes
generally low selectivity and permeability (or, gas flux) through are attractive for such high pressure applications. On the other hand,
microporous inorganic membranes have been a persistent problem that capacity and selectivity of these materials for CO2 are adversely affected

34
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

by increasing temperature, which poses a challenge for practical


applications.

5.3.2.2. Silica-based membranes. Porous silica, made generally by using


sol-gel or hydrothermal synthesis techniques, is often employed not as
the active membrane but as the inert or less active support material for
the more active and CO2 selective membrane or sorption material
deposited on the silica. Accordingly, most of the silica-based meso­
porous membranes used for active separation have either undergone
amine functionalization or grafting on its surface, or impregnation into
the pores in order to improve their CO2 adsorption and separation
properties. At high temperatures (>200 ◦ C) mesoporous silica mem­
branes offer selective separation of H2 from H2/CO2 mixtures by mo­
lecular sieving mechanism, and good permeability that can compete
favorably with the impervious Pd-based membranes commercially
employed in the industry for hydrogen separation. Similarly, amine-
functionalized porous silica membranes at room temperature exhibit
good selectivity up to 800 for CO2 separation from CO2/N2 mixtures due Fig. 41. Relative permeability and selectivity values for a 50:50 CO2:N2
to the preferential adsorption of CO2 [283]. However, typically low CO2 mixture through a NaY zeolite membrane as a function of temperature under
permeability values, which can further be reduced significantly by the dry (open symbols) and moist (closed symbols) conditions. Reprinted from [89]
presence of water vapor in the gas mixture, render silica-based mem­ with permission from the American Chemical Society.
branes fall short of competing favorably with other candidates [89].

5.3.2.3. Zeolite-based membranes. Preferential adsorption and size and zeolite-based membrane studies have been concerned with CO2/CH4
shape selective properties of zeolites are already discussed in Section separation for sweetening of natural gas and CO2/H2 for H2 recovery
5.2.3.1. above and a representative microstructure of a porous alumina- from water gas shift process, the highest values for CO2 separation fac­
supported zeolite membrane is given in Fig. 38. Indeed, zeolite-based tors of 70 and 200 are reported for FAU (faujasite) and T-type (Linde
membranes are usually fabricated on a porous substrate such as stain­ type T) zeolites, respectively, from a 50:50 mixture of CO2 and N2 at
less steel or alumina by hydrothermal synthesis. Size selection (or, room temperature [89].
exclusion) property of zeolites is illustrated in Fig. 40, where the For most types of zeolites, the heats of adsorption of gasses follow the
permeation rates of various gasses including CO2 through two zeolite- trend CO2 > N2 > CH4 > H2 consistent with the strength of electrostatic
based framework structure membranes, namely, ZIF-7 and SIM-1 are interaction of each molecule with the zeolite lattice. As CO2 is smaller in
provided as a function of their kinetic diameters. Because of their similar size than N2 and also adsorbs more strongly on most zeolites, it per­
kinetic diameters and physical properties (see Table 11), the permeation meates preferentially in CO2/N2 and CO2/CH4 mixtures at low tem­
rates for CO2 and N2 are quite similar in both zeolite structures, high­ peratures. At high temperatures, however, competitive adsorption
lighting the difficulty in separating CO2 from N2-rich flue gasses. To diminishes, and results in CO2 selectivity values comparable with its
improve CO2 selectivity, chemical and structural modification of the counterpart in CO2/N2 or CO2/CH4 systems. This is depicted in Fig. 41,
zeolite cages with appropriate cations or amine-functionalization is which also highlights the adverse impact of water vapor in on both the
required to achieve preferential adsorption. Although much of the permeability and selectivity of CO2 near ambient temperatures. How­
ever, the impact of moisture diminishes with increasing temperature, as
does CO2 selectivity [89]. As in almost all solid sorbents and mem­
branes, weakening of the bond with increasing temperature between the
adsorbed species and the surface site is a common drawback that
adversely limits the practical operating temperature regime for these
membranes.

5.3.2.4. Alumina-based membranes. Because of its high chemical, ther­


mal and hydrothermal stability, alumina is a desirable material espe­
cially for high temperature applications. However in most cases,
alumina is employed as the porous support to provide mechanical,
thermal and chemical integrity for the active membrane material. In
those cases where alumina is the active membrane material, it is syn­
thesized as a mesoporous structure with pore sizes in the nanometer
rage, such that the dominant gas transport mechanism is by Knudsen
diffusion. However, alumina has poor selectivity for CO2, and this is one
of the reasons that it is employed more often as a support material rather
than the active membrane.

5.4. CO2 capture by cryogenic distillation


Fig. 40. Comparative permeation rates of pure gasses through two types of
zeolite-base framework structure membranes ZIF-7 and SIM-1 as a function of This is a low temperature capture process that requires cooling, or
their kinetic diameters. Reprinted from [89] with permission from the Amer­ rather refrigeration, of the flue gas down to below the CO2 sublimation
ican Chemical Society. temperature of − 100 to − 135 ◦ C and from 10 to 20 MPa pressure in
order to solidify the CO2 in the flue gas mixture. This cryogenic distil­
lation separates the gasses with respect to their boiling points, and can

35
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

recover up to 90% to 95% of the CO2 in the feed flue stream. As ex­ requires an additional separation step to recover CO2. To avoid this
pected, however, this is a highly energy intensive process that requires additional step, it was proposed that H2 supplied to the anode provides
cryogenic temperatures and high pressures. It is estimated that the en­ the necessary downhill driving force for transporting the carbonate ions
ergy penalty for cryogenic distillation may be in the range of 600 to to the anode [288]. In other words, this configuration schematically
660 kWh/tCO2 recovered in liquid form [284]. depicted in Fig. 42 allows CO2 separation to be accomplished sponta­
neously in a fuel cell mode simultaneously generating electrical power,
5.5. CO2 capture by electrochemical separation rather than in electrolysis mode that would otherwise require expending
external power. An added benefit of this approach is also the fact that
Binary or ternary eutectic mixtures of alkali metal carbonates such as CO2 and H2O are the primary products of the anode reaction, and as O2
Li2CO3, Na2CO3, and K2CO3 are excellent electrolytes that selectively is not a byproduct anymore in this configuration, removing the water
transport CO3− 2 ions through the molten carbonate bath at elevated vapor via condensation from the anode stream allows easy capture of
temperatures. For this reason, alkali metal carbonate eutectics have CO2. However, the reported CO2 permeance (= permeability/membrane
been employed as selective electrolytes in molten carbonate fuel cells thickness) for this configuration was low, less than 5 × 10− 9 mol/(m2.
(MCFC) for many decades [285,286]. MCFC’s typically operate at Pa.s) at 450 to 650 ◦ C with a modest selectivity value of 16 for CO2/N2
temperatures between 500 and 800 ◦ C. This temperature window is well mixtures [288]. The major drawback of this scheme is that the H2 fuel
matched with the typical exit temperature regimes of flue streams from used to drive CO2 separation does not occur in nature, and must be
point source plants and processes. Furthermore, molten carbonate produced first. Accordingly, a life cycle analysis including the produc­
membranes offer thermal and chemical compatibility with CO2-con­ tion of H2 is required to assess the overall energy efficiency and cost of
taining flue gasses for highly selective separation and capture of CO2. separation using this electrochemical approach.
This option eliminates the need to cool down the flue stream to make it Dual-phase membrane systems based on molten carbonates have also
otherwise compatible with the separation media such as been proposed and explored for CO2 separation [289–291]. Variants of
amine-solutions, or membranes based on polymers, MOFs or zeolites. this approach can be grouped under two categories, namely, (a)
The latter group of membrane materials cannot withstand elevated self-driven membranes made typically of a composite of molten car­
temperatures without undergoing irreversible change. bonate and a chemically stable but electronically conducting porous
The general concept of CO2 capture using a molten carbonate fuel support such as stainless steel, and (b) electrochemically-driven mem­
cell configuration goes back several decades [287]. Operationally, the branes, which can either be solely made up of the molten carbonate
molten carbonate membrane configuration for electrochemical separa­ eutectic, or as a composite of an oxide ion conducting but electronically
tion of CO2 may be employed in two different modes. It can either be insulating porous ceramic such as stabilized zirconia that contains the
driven externally by a power source, as in typical electrolysis mode, or molten carbonate within its open porosity.
via the electromotive force generated by the presence of a fuel such as H2 A dual-phase membrane study consisting of a molten Li-Na-K car­
at the anode side, as in a fuel cell mode. In either case, typical anode and bonate eutectic infiltrated into the mixed electronic and oxide-ion
cathode materials are Ni and NiO, respectively. But regardless of the conducting (MIEC) porous ceramic substrate La0.6Sr0.4Co0.8Fe0.2O3
separation mode, membrane configuration or cell architecture, all these (LSCF), reported CO2 permeance values of 2.01 to 4.77×10− 8 mol/(m2.
approaches employ the same reduction reaction of CO2 to carbonate ion Pa.s) at 900 ◦ C for membrane thicknesses between 3.0 and 0.375 mm
at the cathode expressed by, [292]. Using a variant of the self-driven membrane configuration,
another study employed a dual ion-transport architecture, where the
(19) CO3− 2-ion conducting ternary Li-Na-K carbonate melt is coupled to and
2
CO2 + ½ O2 + 2e− = CO3−
2− works in tandem with porous oxide-ion conducting ceramics such as
The carbonate ion, CO3 , is transported through the molten car­
stabilized zirconia and doped-ceria, and reported that these composite
bonate electrolyte and evolves at the anode as CO2 by the reverse of
membranes exhibit rather low CO2 permeances up to 10− 11 mol/(m2.Pa.
reaction (19). Unfortunately, the anode reaction also involves the evo­
s) at 850 ◦ C [290]. Similarly, a ternary Li-Na-K carbonate-based elec­
lution of one molecule of O2 for every two molecules of CO2, and hence,
trolysis concept was demonstrated for capturing CO2, but taking the
concept one step further to electrochemically splitting the CO2 fully into
pure oxygen and high surface area carbon particles [293]. In this study,
it was assumed that the electrical energy required to drive the CO2
electrolysis reaction to carbon and oxygen is supplied by renewable
sources.
Despite the promising features of molten carbonate systems, these
approaches all suffer inherent problems and difficulties, namely, their
highly corrosive nature (i.e., hot corrosion), the dissolution of the metal
support of the electrode materials in the molten eutectic, and their
chemical stability, which all give rise to performance degradation after
long use. In the electrolysis mode, O2 must be separated to obtain a pure
CO2 stream, while in fuel cell mode, energy penalty and cost of upstream
H2 production must be considered as integral part of the assessment
analyses. Also, all schemes exhibit modest values for CO2 permeability,
and hence separation rates, at best. Many of these challenges are
fundamental in nature and difficult to avoid or overcome, which limit
Fig. 42. Schematic of a molten carbonate fuel cell for post-combustion capture the prospects of this interesting electrochemical approach to CO2
of CO2. The flue gas stream is supplied to the cathode compartment where CO2 capture.
undergoes reduction. The presence of the H2 fuel at the anode provides the
driving force to transport the carbonate ion in the molten carbonate eutectic,
while the electrons traverse through the external circuit from the anode to the 6. Direct capture of CO2 from ambient air
cathode performing useful electrical work Reprinted from [89] with permission
from the American Chemical Society. One of the ways to achieve negative emissions is to capture CO2
directly from the atmosphere. Although methods of removing CO2 from
air in enclosed environments such as submarines go back for decades,

36
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

the concept of capturing CO2 directly from ambient atmosphere for the domain of post-combustion separation. Most DAC processes under
explicit purpose of reducing atmospheric concentrations was first development involve scrubbing large volumes of air in alkaline solutions
advocated nearly two decades ago [294]. Despite concerns raised and capturing CO2 as solid carbonates, which are then calcined at
regarding the economic and technical feasibility of direct air capture elevated temperatures to recover the CO2 in highly pure form [103].
(DAC), there has been growing interest and activity to explore this To appreciate the scale of the challenge to DAC or CO2 capture in
negative emissions technology [104,295,296]. Proponents of DAC have general, note that to reduce atmospheric concentration of CO2 by only 1
long argued that even if carbon emissions were to be stopped all together ppmv, it is necessary to capture and remove nearly 7.8 Gt of CO2 from
today, CO2 that has already accumulated in the atmosphere cannot the atmosphere [45]. This is a daunting task, and unfortunately, we do
eliminate the risks of climate change and incipient warming. So atmo­ not yet have cost-effective CO2 capture technologies available at the
spheric CO2 concentrations need to be reduced expeditiously. It is multi-gigaton scale that can be deployed to mitigate emissions and affect
further argued that climate change mitigation strategies should involve atmospheric concentrations.
a broad range of diverse technologies that include direct air capture, in As expected, the low concentration of CO2 in the atmosphere renders
order to affect a net reduction in the current atmospheric levels. How­ DAC an economically costly process. For DAC to compete with other CO2
ever, CO2 capture requires large amounts of water, and hence must be capture alternatives, it needs to become cost effective at large scale. The
sited in close proximity to a water source. Moreover, DAC is not site cost of DAC was estimated to be $610/tCO2 excluding compression for
specific and does not need to be located near emission sources. It offers transportation and sequestration, based on the average carbon intensity
the important opportunity to capture CO2 emissions from distributed (i.e., carbon footprint per unit of energy produced) of 610 kgCO2/MWh
sources, which make up nearly ½ of the global CO2 emissions, as for the US power grid in 2011 [45]. Comparing this to the reference
opposed to most of the other technologies, which cater more to 2011 cost of ~$80/tCO2 for capturing from coal power plant flue gas
capturing from point emission sources. This is an important distinction stream, DAC was deemed prohibitively expensive. More recently, a
and an advantage for DAC. 2016 U.S.-DOE task force report suggested that costs must be brought
Most of the CO2 capture technologies, processes and materials down to below $200/tCO2 by technological advances for DAC to offer
reviewed in this article above are also relevant and applicable to DAC, cautious viability [31]. The report also concluded that DAC might play
and will not be discussed here further. Only the salient features of DAC at best a modest role in climate change mitigation over the next several
are reviewed and assessed. Recent progress and advances in sorbent decades, and only if based on renewable sources. Another study esti­
materials and capture processes for DAC are recently reviewed exten­ mated the cost of cutting CO2 emissions by replacing carbon-intensive
sively elsewhere [297]. Besides technological CCS processes, there are of power generation technologies with solar cells to be around
course naturally occurring terrestrial mechanisms that help capture CO2 $500/tCO2, challenging the investments made for solar cells [300].
directly from the atmosphere including uptake by large bodies of water Others questioned if air capture is the best way to spend resources,
such as lakes and oceans, biomass growth on land and in sea, and which may take funds away from developing more cost-effective solu­
mineralization into carbonates by naturally occurring sediments and tions, and concluded that DAC may make sense only after all CO2
minerals on land. However, these processes are generally considered emissions from centralized and point sources are captured [45].
outside of DAC’s domain of interest, and not covered here. A recent techno-economic study estimated the levelized cost of DAC
While the atmospheric concentration of CO2 (~ 414 ppmv) is to range between $94 and $232/tCO2 when CO2 is delivered at 15 MPa
alarmingly high for climate mitigation but quite low for effective cap­ using either 8.81 GJ of natural gas, or a combination of 5.25 GJ of
ture, it is not surprising that DAC is a highly energy intensive and costly natural gas and 366 kWh of electricity, per ton of CO2 captured from air
technology. Indeed, the theoretical minimum energy requirement for [301]. These values highlight the high thermodynamic barrier for CO2
DAC is ~22 kJ/molCO2 at room temperature. Naturally, the actual separation from air that requires significant energy expenditures for
amount of energy demand for DAC is several times larger. Indeed, a capture, and likely to create new emissions unless this energy is supplied
recent techno-economic modeling study assuming a scaled up DAC ca­ from renewable sources. Indeed, it was suggested that DAC would be
pacity increasing by 1.5 GtCO2/yr has estimated the energy requirement incompatible with high carbon energy sources such as fossil fuels [45].
for DAC to be up to 300 EJ/yr (300×1018 J/yr) by 2100 [298]. DAC The CO2 separation step in DAC is energetically the most intensive and
technology also requires a large real estate footprint, as it relies on large financially the most expensive step. Indeed, the minimum theoretical
contact areas for effective CO2 uptake during the scrubbing or absorp­ work requirement is more than 20 kJ/molCO2 [103], and the mechan­
tion operations. This necessitates huge contactor facilities with nearly ical work to isothermally compress the recovered CO2 from 0.1 to 6 MPa
300 times more contactor area than would be required to capture CO2 pressure adds another 11 kJ/molCO2 [104]. Accordingly for any carbon
from a combustion point source such as the exhaust stream of a power capture and sequestration (CCS) technology to be viable, the down­
plant [299]. Indeed, large volumes of air need to be processed to over­ stream benefits of capturing CO2 from air should exceed these minimum
come the unfavorable CO2/air ratio of ~1/2500 and to remove mean­ energy requirements.
ingful quantities of CO2. Its low concentration in air also discourages Despite the economic and technical risks, several start-up companies
many of the technologies that particularly require heating, cooling, and are already exploring DAC technology for commercial purposes, as part
compression from consideration. Moreover, similar molecular diameters of a complementary solution to help reduce anthropogenic CO2 in the
of CO2 (0.330 nm), N2 (0.364 nm) and O2 (0.346 nm) render atmosphere. Notably, Carbon Engineering in Canada and Climeworks in
size-selective separation of CO2 by membranes practically ineffective. Switzerland have been developing DAC technologies, and there are now
This leaves absorption or adsorption as technically and economically about 15 DAC plants already operating in Europe, the United State and
more viable choices for DAC. These sorption technologies generally Canada [302]. However, their capacities are all rather small. For
involve chemical sorbents that exhibit high binding affinity for CO2 example, the largest DAC plant so far, named Orca, currently under
relative to N2 and O2. These chemical sorbents include aqueous alkaline construction by Climeworks and planned to start operations in 2021
solutions of suitable sorbent hydroxides such NaOH, KOH, and Ca(OH)2, aims to capture merely 4000 tCO2 annually [302]. Such small capacities
hybrid organic-inorganic sorbents, or a rich variety of amine-based also help highlight the enormous challenges of capturing at the gigaton
sorbents often supported on solids to capture CO2. Aqueous solutions scale that would be required to make a difference.
offer the advantage of longer air-water contact times, as well as cheaper
and continuous operation of cooling towers. Fortunately, many of the 7. CO2 transport
technological options discussed above for post-combustion separation
including solvents, membranes and sorbents are also being considered Commercial transport of carbon dioxide may be accomplished by
and assessed for DAC, as DAC can in principle be viewed within the pipelines, tanks, rail, and ships [303–307]. For distances typically of the

37
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

order of 1000 km, pipelines are the most common and commercially depend on the volume of CO2, the length of the pipeline, and the dis­
mature way to transport CO2 from point emission sources to the point of tance and location of capture site with respect to the storage site. For
use, usually for enhanced oil recovery (EOR) and other applications. For example, it was estimated that depending on the location of the emission
distances much longer, or across large bodies of water, transport by ship point source, the costs for pipeline transport might vary by 30%.
is more preferable. Rail or tanker transport may be suitable for small Moreover, for a 500 MW plant located in the Midwest of U.S.A., it was
quantities CO2 over short distances. estimated that CO2 transport costs vary widely from $0.15/tCO2 for a
Depending on the method used for capturing it, CO2 may contain 10 km long pipeline to $4.06/tCO2 for a 200 km pipeline [309]. Simi­
impurities such as SOx, NOx, amines, NH3, CO, H2, N2, and O2, besides larly, an analysis in China estimated the transportation costs for 4000
moisture. These impurities affect the physicochemical and hydrody­ ton CO2 per day (or, 1.46 million tons/year) to be $12.64/tCO2 for
namic properties of the CO2 gas mixture, including its viscosity, railroad tankers, $7.48/tCO2 for ship tankers, and $7.05/tCO2 for a
compressibility, and fluid dynamics, as well as the corrosion and 300 km pipeline [310].
potentially hydrogen embrittlement of the transport vessel or pipeline. CO2 transport by ship is an alternative solution to pipeline transport
Hence, these impurities may have to be removed and the CO2 gas especially to offshore enhanced oil recovery sites, and involves
dehumidified prior to compression and transport. Moisture is a critical employing large volume (~20,000 m3) pressurized vessels or tankers at
component and demands special consideration. Dry CO2 typically con­ or above the triple point of CO2 to achieve high storage densities. A
taining < 10 ppmv H2O is not corrosive to pipelines made of carbon- North Sea tanker transport study of CO2 indicated that the liquefaction
manganese steels that may also tolerate the presence of other contami­ process constitutes ~77% of the total energy requirement of 142 kWh/
nants. On the other hand, moist CO2 is highly corrosive even at tCO2 for the entire transport chain, and the cost of ship transport was
moderately high temperatures and requires highly expensive corrosion estimated to be $20 to $30/tCO2 for annual shipping volumes larger
resistant alloys that are internally coated with an additional corrosion than 2 MtCO2 and for distances limited to the North Sea [306]. Clearly,
barrier layer [308]. Hence, moisture is generally removed from CO2 the North Sea tanker transport study seems to indicate shipping costs are
before transportation in order to avoid the high cost of pipeline material. more expensive than the cost of pipeline transport, and highlights the
For distances significantly longer than 1000 km, sea transport offer importance of the mode of transport as well as the regional and
better economics and is generally preferred, but nevertheless require geographic economics.
similar measures of removing moisture. For long distance transport of CO2 over land, pipeline transport is
CO2 can be transported in several forms, including gas, liquid, or as still considered the most viable option, especially if the emission source
supercritical fluid. The latter can be achieved when CO2 is compressed is a point source such as a power plant with an economic lifetime
above the critical pressure of 7.38 MPa at or above 31 ◦ C as shown in generally of several decades, which is compatible with the term of in­
Fig. 43, where the densities of the liquid and gaseous CO2 phases become vestment for building the pipeline. However, for short distances be­
equal to each other and behave as a single fluid. However, the tween the CO2 source and storage, or for shorter source lifetimes, road,
compressibility of CO2 in the typical operating regime for pipeline railroad or even ship tankers may be more cost-effective.
transport is nonlinear, and sensitive to the presence of impurities such as Pipeline transport is a mature technology and supercritical CO2 is the
H2S [309]. Operating the pipeline at higher pressures (>8.6 MPa) re­ preferred form. This requires that the pipeline must be operated above
duces the nonlinear effect and increasingly pulls it more into the linear the supercritical pressure and temperature. Usually, supercritical CO2
regime. pipelines are operated at pressures between 8.5 and 15 MPa, and tem­
In order to maximize the mass-to-volume ratio during transport, peratures between 13 ◦ C and 44 ◦ C in order to maintain the stable single
dense phases, namely, liquid or supercritical CO2 are preferred. phase flow of CO2 [311]. Moreover, large diameter pipelines lower the
Expectedly, there is an energy penalty to achieve this state, because the pressure drop and the associated pumping losses, and hence may be
transport vehicle or the pipeline must maintain at all times the pressure more energy efficient. But they also cost more to built. So the optimum
and temperature conditions commensurate with the transported phase design must minimize both energy penalty and cost of transport. Some of
requirements as either liquid or supercritical CO2. the key considerations for pipeline transport of CO2 include the pipeline
The CO2 transport system or mode chosen between the capture and temperature and pressure, hydrate formation, corrosion, CO2 impurities,
storage sites must be safe, reliable, energy efficient, environmentally and last but not least, the safety and reliability to avoid catastrophic
friendly, and cost-effective. The choice of transport mode, whether it is failure. In Northern America, there is already 6560 km of installed CO2
by tanker ships and trucks, railroad tankers, or by pipeline, may also pipeline capacity to annually transport 150 million tons of CO2 for
enhanced oil recovery (EOR) purposes. However, this is far from the 37,
000 km of CO2 pipelines needed in the U.S. only, and over 150,000 km
needed in the European Union for large scale deployment of CCS by the
year 2050 [187].

8. CO2 storage options

Large-scale CCS at the multi-gigaton scale is considered as an effec­


tive way of mitigating CO2 emissions and climate change [187,188,312,
313]. Onshore or offshore deep underground storage of CO2 may benefit
from technologies and knowhow developed previously by oil and gas
industries. Geological structures and cavities, abandoned or depleted oil
and gas reservoirs, depleted coal mines, and saline aquifers may be
suitable underground options also for long term CO2 storage provided
that they offer effective trapping mechanisms to store large quantities in
a safe and economic manner. This requires fundamental understanding
and assessment of cooperative mechanisms and processes pertaining to
Fig. 43. CO2 phase diagram, depicting the critical point. Reprinted from [133] the hydrology, geochemistry, and geomechanics of the storage sites
with permission from Elsevier. [314]. Injecting vast volumes of pressurized CO2 into geological struc­
tures and formations leads to adverse effects including increased sub­
surface pressure, induced seismic activity, contaminated aquifers and

38
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 44. Depth dependence of CO2 density, where the cubes depict the relative volume of CO2 with rapid contraction under hydrostatic pressure until about 800 m,
where CO2 starts to display supercritical behavior [312].

subsurface water, and compromised storage integrity [315]. There are area, as CO2 is lighter than air and settles down close to the ground.
also socioeconomic, environmental, geological, financial, regulatory, Similarly, gradual leakage may be due to faults or fractures in the in­
security, and policy considerations in the selection and evaluation of jection well or the well cap, or geological fissures. Susceptibility to
storage sites. The multi-dimensionality of these concerns makes it seismic activity either naturally occurring or activated by large-scale
challenging to devise consistent feasibility protocols for the assessment geological storage [315] must also be taken under consideration as
and qualification of potential reservoirs for long-term CO2 real threats to the integrity of the seal that may trigger, when compro­
sequestration. mised, abrupt or gradual leakage of CO2.
Incidentally at depths below 800 to 1000 m, supercritical CO2 pos­ It is clear many of the storage options face multiplicity of technical,
sesses a density comparable to liquid (see Fig. 44) and can be trapped economic, geographical, social and regulatory hurdles. For CO2 capture
effectively underground. Some of the effective storage mechanisms to have a significant impact on reducing emissions, it has to be deployed,
include trapping inside the pores of sedimentary rocks or below an implemented and stored at the multi-gigaton scale annually. Fortu­
impervious rock layer (i.e., caprock), adsorption on organic matter in nately, there is ample potential capacity for CO2 storage across the
coal mines or shale structures, dissolution in underground fluids or globe, with sedimentary basins offering 5,000 to 25,000 GtCO2, oil and
aquifers, and immobilization by reaction with natural minerals to form gas reservoirs around 1000 GtCO2, and saline aquifers adding 4,000 to
solid carbonates (i.e., mineralization). Similarly, offshore underground 23,000 GtCO2 [187]. The geographic distribution of potential CO2
storage deep under the ocean is also possible. In fact, the Sleipner CCS storage reservoirs is illustrated in Fig. 45. It was suggested that under­
project, operational since the mid 1990′ s, involves offshore storage of ground storage in deep saline aquifers potentially offers long lifetime for
CO2 in an underground saline formation beneath the North Sea bed. CO2 storage exceeding 100 years [316]. At CO2 storage densities of
However, for all onshore or offshore storage options to be viable, CO2 nearly 0.6 kg/m3 [71], storing the current annual CO2 emissions that
leakage rates from storage sites should be negligibly small, or there is exceed 37 Gigatons would require nearly 62,000 km3 of storage volume.
onsite recovery capability to capture the leaked CO2. Two major types of Such massive emission quantities dwarf the current industrial produc­
leakage may occur at the storage site. Abrupt leakage may be caused by tion capacity worldwide (see Table 3), and pose further technical and
catastrophic failure of the injection well or the well cap, which could financial barriers for CCS in terms of the size scale of the problem.
pose great health and environmental hazard and even loss of life to CO2 sequestration for long term storage presents additional chal­
plants, animals and humans especially in the low lying surrounding lenges associated with safety, cost, proper site selection, induced seis­
micity [315], CO2 leakage, and active monitoring of both the storage
cavity underground and the surface site. The 2015 National Research
Council report [32] recommends further investigation of these concerns
posed by CO2 storage to fully understand and anticipate the potential
risks. It also recommends a portfolio of actions for developing advanced
CO2 removal technologies that offer the potential of reducing net
emissions. It recognizes that low risk strategies and technologies
currently available are limited by cost and fall short of removing
climatically significant quantities of CO2 beyond the natural processes
by land and sea.
Despite immense progress in renewable energy technologies
including solar and wind, their share in the total energy production
remains a growing but yet a small fraction, and lack of cost-effective,
utility scale, long-term energy storage technologies also limit their
Fig. 45. Geographic distribution of potential CO2 storage capacities [187]. rapid deployment [101]. It is likely that fossil fuels continue to play a
Reproduced with permission from the Annual Review of Environment declining albeit a major role in the global energy economy into the
and Resources. foreseeable future [22]. Accordingly, sequestration and long term

39
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 46. Comparative locations of large-scale CO2 emission sources (top) and potential CO2 sequestration sites (bottom), indicating good correlation and reasonable
proximity between the sources and storage sites [312].

storage of CO2 emissions is considered the most viable option to stabilize void fraction of mineable and unmineable coal beds also offer the op­
the atmospheric CO2 concentration in the coming decades. An earlier portunity to extract the methane that is trapped in the pores of the coal
study estimated the total carbon storage needed in the 21st century to be seams. Indeed, methane extraction from coal beds by CO2 has been
between 600 and 2400 Gt of carbon (or, 2200 to 8800 Gt of CO2) [188]. successfully used in practice [320]. However, despite their adequate
More recently, the practically accessible global capacity for geologic capacity, coal seams usually have low permeability. Furthermore, esti­
sequestration of CO2 is estimated to range between 8000 and 55,000 mation of their potential storage capacity, storage integrity and reli­
GtCO2 [317]. Clearly, there is sufficient geological storage potential for ability, and environmental and safety concerns pose serious challenges.
CO2 sequestration. For large-scale CO2 storage, the proximity of the Accordingly, coal bed storage is generally viewed as a challenging
sequestration site to point sources is also an important consideration to option.
reduce the cost, logistic and technical challenges of CO2 transportation. Besides storage capacity and proximity to emission source, CCS
Fortunately, there is good matching between the geographical distri­ scenarios also require consideration of the storage times and leakage
bution of large-scale CO2 emitting sources around the globe and po­ potential. The natural sequestration process of mineralization into car­
tential terrestrial storage sites [312], as depicted in Fig. 46. A similar bonates provides long time frames for storage (of the order of 105 years)
correlation and proximity exists between fossil fuel power plants and and vast capacities approaching 105 GtCO2, albeit at slow uptake rates
potential geological storage sites in Northern America [318]. [188]. Similarly, CO2 storage by absorption in oceans as carbonic acid
Regional capacities for CO2 storage have also been assessed by may look attractive, but it trades one environmental problem with
governmental and scientific organizations in the last decade [319]. The another as acidification of the oceans by CO2 uptake acutely threatens
results of this study indicated that there is ample capacity for CO2 the marine ecosystem and coral life. Likewise, sequestration in biomass
storage around the world, with North America, Brazil and China offering such as afforestation and reforestation may be limited by growth rate,
the largest capacities far exceeding 1000 GtCO2. However, note that time scale and climatic constraints. So it is generally believed that
these global estimates reflect total available sizes, but the actual sequestration in underground structures, formations and aquifers may
accessible or usable capacities that are economically and technically provide easier and more expedient route for long term storage of CO2. It
feasible may be much less. With respect to source types, storage capacity is also proposed [188] that perhaps another reliable way to store CO2
estimates for sedimentary basins range between 5,000 and 25,000 Gt of permanently is by carbonization, where CO2 is first injected into
CO2, saline aquifers can similarly accommodate 4,000 to 23,000 Gt of geological formations, such as basalt, that typically contain
CO2, and reservoirs for oil and petroleum, after full depletion, provide alkaline-based reactive minerals. The CO2 dissolves in the water trapped
about an additional capacity for 1000 Gt of CO2 [187]. Storage in the in the pores of the rock formations. The acidified water leaches out the

40
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Table 16
Summary of some major geological CO2 storage projects worldwide [322].
Project Operational Since Stored Amount Storage type
(MtCO2/yr)

Sleipner Oct. 1996 0.9 Offshore aquifer


Snøhvit April 2008 0.7 Offshore aquifer
Quest Nov. 2015 1.0 Onshore aquifer
Illinois Industrial CCS April 2017 1.0 Onshore aquifer
Tomakomai April 2016 0.1 Offshore aquifer
Gorgon Autumn 2017 3.4 Offshore aquifer

minerals in the rock by reacting to form carbonates and bicarbonates,


whereby generating stable products that eliminate concerns related to
long-term leakage [321].
The first commercial CCS project that was undertaken for climate
mitigation was the Sleipner field in the North Sea 250 km off the shore of
Norway that is still in operation since 1996. Another early CCS project In
Salah that started about a decade ago in Algeria stopped its operations
recently. Earlier projects between the 1970′ s and 1990′ s provided
reduction in CO2 emissions but did not target CCS specifically for
climate mitigation purposes. There are several CCS projects already in
operation (see Table 16), and many industrial scale CCS projects are
under various stages of development or construction in the U.S., Nor­
way, Canada, and Australia. Unfortunately, the stored CO2 vol (~ 1
MtCO2/yr) in these projects are clearly several orders of magnitude
lower than the annual emissions of nearly 37 GtCO2. Recent progress has
increased the global capacity for carbon capture by nearly 33%,
including the addition of 12 new commercial projects in North America
Fig. 47. Impact of increasing CO2 content (top) on ocean acidity (bottom)
in 2020. Currently, there are 65 commercial CCS facilities under various monitored between 1988 and 2007 [328].
stages of development and completion around the world, and 26 that are
already in operation capture nearly a total 40 MtCO2 annually [323,
poses some risks. Increased CO2 content alters ocean chemistry and the
324]. Again, this annual capacity falls far too short of making a dent in
marine ecosystem. Dissolved CO2 increases ocean acidity (see Fig. 47),
mitigating the continued emissions of CO2 into the atmosphere.
which is implicated in the destruction of coral life and marine habitat.
The chain costs associated with CO2 sequestration are uncertain and
Unlike geological storage, however, there are no plausible and
depend on many factors. Nevertheless, an early estimate had projected
abrupt mechanisms for ocean storage that can cause catastrophic release
that ~$30 per ton of CO2 may be achievable in the long term [188]. It
of CO2 into the atmosphere. On the other hand, gradual release may
was also argued then that development of cost effective and reliable CCS
certainly occur, partly due to ocean warming and also by mixing and
technologies would present a cheaper option than a rapid and complete
circulation by currents over centuries [329]. As sea temperatures rise
transition from fossil fuel-based power generation to carbon-free
due to atmospheric heating, ocean’s buffer capacity for CO2 uptake will
renewable technologies. As expected, the cost of CCS depends on
decrease due to the reduced solubility of CO2 in warmer seawater. Many
many factors. For example, the total cost of CCS from coal-based inte­
of these outstanding issues and knowledge gaps require fundamental
grated gasification combined cycle (IGCC) plant is estimated to be
understanding for the prospects of long term large-scale storage of CO2
around $4.5/tCO2 with the end use targeted for enhanced oil recovery,
at the ocean floor to gain comparable level of interest and momentum as
while CCS from a natural gas combined cycle plant with long-term
geological sequestration.
storage of CO2 in saline formations may cost significantly much higher
at $72.4/tCO2 [92]. Variations in the cost of CCS are also valid for other
9. Downstream utilization and transformation of CO2
industrial processes such as iron and steel, cement, refinery, and pulp
and paper, where cost estimates range between $30 and $70/tCO2
Although there is ample geological and deep sea capacity for CO2
depending on the industry [41].
storage as discussed in the previous section, the option of re-utilizing
Besides geological structures and underground formations, storing
CO2, at least in part, in industrial applications, or transforming it into
CO2 deep on the ocean floor is another alternative for large-scale storage
useful chemicals and fuels [36,38,330–335] certainly warrants due
[312]. Oceans offer enormous storing capacity and have absorbed more
consideration as it helps reduce the amount that would otherwise be
than 600 GtCO2, or nearly 40%, of the anthropogenic CO2 emissions
emitted into the atmosphere. CO2 is a valuable source for carbon that
since the industrial revolution [325]. They cover nearly 70% of the
forms the backbone of wide range of indispensible industries and
world’s surface with an average depth of about 3800 m, and have served
products including polymers and plastics, petrochemicals, pharmaceu­
as a major sink for removing CO2 from the atmosphere with an average
ticals, food packaging, beverage industries, refrigeration, fire extin­
annual uptake rate of nearly 7 GtCO2.
guishers, synthetic fibers, and chemicals and fuels. As long as carbon
CO2 sequestration in the ocean by reacting power plant flue gasses
remains chemically bound in these products and materials during their
with seawater followed by carbonate dissolution has been proposed
life cycles, it is effectively blocked from entering the atmosphere and
earlier to accelerate the geological time scales for uptake and also to
hence considered ‘chemically stored’. However, carbon life cycle con­
minimize the impact of otherwise increased acidity in the ocean [326,
siderations should include all upstream as well as downstream pro­
327]. Alternatively, captured CO2 can be injected deep into the ocean
cesses, and not only focus on the final product. Although CO2 utilization
floor to depths of about 3000 m, where CO2 density exceeds that of
is not generally viewed as a game-changing strategy for emission miti­
seawater and sinks to the bottom. Injected CO2 can form solid CO2 hy­
gation, it nevertheless helps recycle the carbon to produce useful
drates or CO2 lakes on the seabed, and can thus be immobilized possibly
products, especially if the energy input is supplied from a renewable
for centuries. However, this technology is not well understood yet, and

41
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

source. Dwarfed by the vast scale of CO2 emissions [5], we are also Table 17
severely limited by our global industrial capacity and in our ability to Comparison of energy densities of liquid fuels and batteries, indicating the su­
channel and process even a fraction of the CO2 emissions for periority of chemical storage over batteries for long duration energy storage
re-utilization and transformation purposes (see Table 3). Nevertheless, Reproduced from [101].
development of cost effective and energy efficient technologies for CO2 Gravimetric Energy Density Volumetric Energy Density
conversion and utilization as a C1-building block for the production of (Wh/kg) (Wh/L)
chemicals and fuels, even in part, certainly offers the opportunity to Liquid Fuels
close the carbon cycle and help reduce demand for long term storage of Gasoline 12,330 9.060
CO2 [79]. Although not in volumes anywhere near emissions, there are a Diesel 12,700 10,700
Propane 12,870 7,490
variety of existing applications already commercialized that exploit the Butane 12,700 7,190
chemical and physical properties of CO2. Ethanol 7,490 5,890
Methanol 5,620 4,470
Batteries
9.1. Enhanced oil recovery Li-ion battery* 150–210 450
Ni-metal 70 220
Currently, the largest industrial use of CO2 is in enhanced oil re­ hydride*
Zn-air battery* 300 240
covery (OER) applications, where CO2 is injected under pressure into
depleted or semi-depleted oil and gas reservoirs to force out the residual *
Practical values.
fossil fuel trapped in the geological formations. EOR is a mature tech­
nology that has been widely used for many decades, and provides great
economic benefits to the oil and gas industry, but most fields have
annual injection rates of only several MtCO2 [189]. Up to 40% of the
residual fossil fuel can be extracted by EOR, while the CO2 is perma­ The conversion of CO2 into useful fuels accomplishes two objectives
nently trapped and stored in the reservoir. A recent study by the U.S. simultaneously. First, it recycles CO2 and transforms it into a useful
National Energy Technology Laboratory (NETL) based on the possibility product that closes the carbon cycle. If energy from renewable or non-
of cost effective capture of 20 Gt of CO2 from power plants and other carbon sources such as nuclear, solar, wind or hydrothermal is
industrial sources has estimated that nearly 67 billion barrels of oil can employed for this conversion, the process can also be carbon neutral.
be recovered from economically recoverable resources using EOR [336]. Equally important is the fact that due to the high energy density of
Achieving this goal requires major reduction in the cost of carbon hydrocarbons, CO2-to-fuels conversion also offers an effective means for
capture systems to approximately $40/Gt of CO2 captured at the plant large-scale “energy storage” technology that is badly needed for utility-
gate by the year 2025, excluding transportation and storage. Although scale renewable power, extending its efficacy from days to months for
only 47 MtCO2 is annually employed worldwide for OER, a 2011 IEA the electric grid [101]. Especially liquid fuels are convenient and
report estimated that implementation of state-of-the-art OER technology portable energy carriers with high energy density, and are attractive for
in 50 of the world’s largest oil fields could produce 470 billion barrels of a variety of energy production and storage applications. Table 17 sum­
additional oil and store 140 GtCO2 [337]. marizes and compares the energy densities of major liquid fuels with
battery storage and highlights their significant advantage.
CO2-to-fuels conversion can be considered in a similar realm as
9.2. Other industrial applications
power-to-gas, power-to-liquid or power-to-X [335,339–342], and is ex­
pected to play a major role in storing electrical energy in the form of
Aside from enhanced oil recovery (EOR), CO2 has long been
fuels such as hydrogen, methane, ammonia, or methanol, and useful
employed in many industrial scale applications including sodas and
chemicals such as syngas (H2+CO), CO, or formate. While power-to-gas,
other carbonated beverages, in fire extinguishers, as well as refrigera­
power-to-liquid and power-to-X are used interchangeably to indicate
tion. More recent applications exploit its remarkable physical and sol­
that the electrical energy for these conversion processes are supplied
vation properties. Liquid CO2 has very low viscosity and excellent
exclusively by renewable sources, CO2-to-fuels concept embraces all
wetting properties, and a density comparable to many of the other
types of energy inputs. Although the costs of many power-to-gas pro­
commonly used organic solvents. Similarly, supercritical CO2 is an
cesses such as water electrolysis and methanation are predicted to
excellent solvent and employed for many extraction and separation
decrease 75% by mid-century [343], techno-economic and life cycle
processes in the food industry, and cleaning operations for the garment
assessments of many of the power-to-gas technological options indicated
(i.e., dry-cleaning) as well as the microelectronics industries. Both liquid
that they still fall short of economically competing with conventional
and supercritical CO2 are considered “green solvents” employed
gas production processes [38,344]. Feasibility of many power-to-gas or
increasingly in a wide range of industrial applications including
-liquid processes is quite sensitive to the price and availability of
replacing organic solvents for precision coatings in automotive and
renewable power. A recent study concluded that for electrosynthesis of
furniture manufacturing, polymer production and processing, biomed­
fuels from CO2 using renewable power to be able to compete favorably
ical applications, catalysis, and decaffeinating coffee and tea [338].
with fossil fuel-derived synthesis processes, the price of renewable
electricity needs to fall below $0.04/kWh and the electrochemical
9.3. CO2-to-fuels conversion conversion efficiency needs to be > 60% [345].
There are several thermal routes for CO2 re-use for the synthesis of
CO2 capture and utilization (CCU) complements CCS as CO2 is a fuels and chemicals by partially reducing it to CO or C1 hydrocarbons,
valuable source of carbon for chemical feedstocks and also as C1 including thermal catalytic conversion [346,347], solar thermal [348],
building blocks for the production of high value C2 and higher hydro­ and plasma [349,350]. Also, CO2 can be converted to hydrocarbons
carbons and fuels. Interestingly, the global market volume for hydro­ directly by photocatalytic [334,351,352], electrochemical [332,345,
carbon fuels is nearly two orders of magnitude larger than that for 353–355], or photoelectrochemical [356–358] pathways, or indirectly,
chemicals, so CO2-to-fuels is an attractive goal and a qualified propo­ by partial reduction first to CO that is then mixed with hydrogen to
sition. Although this conversion process, depicted by the overall reac­ undergo a catalytic conversion process to produce hydrocarbons via
tion (20), is technically viable, it is highly energy intensive. Fischer-Tropsch (F-T) synthesis. The F-T process, developed in the
1920′ s in Germany and currently employed in Sasol plants in South
CO2 + H2 (or, H2O) = CO, H2, Hydrocarbons (C1, C2, C3,…) (20)

42
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Africa, is an established and mature technology that allows catalytic


conversion of CO-H2 mixtures typically of the CO:H2 ratio of 1:2 into a
wide range of hydrocarbons and liquid fuels [335,359–362].
As CO2 represents the lowest energy state in the C-O binary system, it
is chemically a very stable molecule (ΔGfo = –396 kJ/mol). Any reaction
that involves its reduction is thermodynamically uphill with an energy
barrier of ~ 1 eV. Nevertheless, this energy barrier may be affected and
lowered by catalysis, and more so by electrocatalysis [109]. Indeed by
imposing a suitable potential bias from an external power source, elec­
trochemistry makes it possible to alter the energy profiles of reactions so
much so as to even help reverse the energy barrier for a thermody­
namically uphill reaction to a downhill gradient. This is the underlying
principle of many electrolysis and electrosynthesis reactions, as well as
electrowinning and electroplating processes.
The first demonstration of electrochemical conversion of CO2
directly into methane fuel over transition metal catalytic electrodes
deposited on an oxide-ion conducting solid electrolyte yttria-stabilized
zirconia (YSZ) was reported in the early 1980′ s [363–365]. A slight
variant of this electrochemical approach was proposed for life support Fig. 48. Comparison of energy efficiency and current densities for the elec­
systems that directly dissociated CO2 for oxygen recovery [364–368] or trolytic conversion of CO2 into formic acid, syngas, and hydrocarbons (pri­
simultaneously generated oxygen and methane from CO2/H2O mixtures marily, methane and ethylene) in aqueous electrolyte containing electrolyzers.
using a solid state electrochemical reactor [369,370]. It was also pro­ Efficiencies of the alkaline and PEM electrolyzers reflect total system level ef­
posed that the reaction mechanism for the high temperature electro­ ficiency, while efficiencies for CO2 conversion include only the cathode losses
while neglecting anode and other system losses. Reprinted from [372] with
chemical conversion of CO2 to methane involved surface CO as the
permission from the American Chemical Society.
intermediary, and the rate-determining step was the dissociative hy­
drogenation of the adsorbed CO [363,365]. The results further indicated
that electrochemical stripping of oxygen sequentially from adsorbed
CO2 and CO has enhanced the turnover frequency of the methanation
and energy conversion efficiencies of 25% to 50% [345]. Fig. 48 sum­
reaction by nearly two orders of magnitude at 650 ◦ C and the reaction
marizes the results from eighteen different studies, and maps the prod­
rate constants exhibited strong dependence on the externally applied
uct distribution and conversion efficiencies obtained on various cathode
potential that drove the uphill reaction.
materials [372]. It is clear that these low (or, room) temperature elec­
Interestingly, the generally accepted stepwise reaction mechanism
trochemical methods offer only moderate efficiencies and require large
currently accepted for room temperature electrochemical reduction of
overvoltages to drive the CO2 reduction and oxygen evolution reactions.
CO2 also involves CO as the intermediary [353]. From a practical point
Even at a modest current density of 50 A/m2, the CO2 reduction over­
of view, there are obvious advantages accomplishing CO2 reduction at or
potentials for most metal cathodes are around 1 V, and for Cu cathode it
around room temperature, including the chemical stability of higher
is 0.9 V for the CO2 to CO reaction and 1.2 V for CO2 to methane [353].
hydrocarbons, which are unstable at elevated temperatures, avoiding
Hence, development of improved electrocatalysts with well-defined and
large thermal loads required for high temperature electrolytic reduction
controlled microstructures is needed to reduce electrode overvoltages
of CO2, and employing water-based electrolytes as the H2 source for the
and render this electrochemical approach more suitable for practical
hydrogenation reaction of CO2 to hydrocarbons.
considerations. The low solubility of CO2 in aqueous electrolytes is
Copper has emerged as the most preferred room temperature elec­
another limiting factor. Non-aqueous electrolytes offer significantly
trocatalyst for CO2 reduction, because of its exceptional ability to
higher solubility for CO2 than their aqueous counterparts. For example,
catalyze the conversion of CO2 to wide range of chemicals and hydro­
dimethyl-formamide can dissolve more than 20 times more CO2 than the
carbons. Indeed, since the first demonstration of room temperature
same quantity of aqueous solutions, while propylene carbonate can
electrochemical reduction of CO2 on Cu electrocatalyst [353,371], there
provide 8 times more solubility for CO2 [380].
has been intense activity for low temperature electrochemical routes
The reaction mechanisms for CO2 reduction and conversion in liquid
[353,372–374]. A recent study demonstrated the highest turnover fre­
electrolytes are complex [378], and not well understood. The product
quency of 4.3 s − 1 and 1.8 s − 1 (corresponding to partial current den­
distribution, or selectivity, and reaction pathways depend upon many
sities of 132 and 84 A/m2) respectively for the electrochemical
process variables, and strongly on catalytic properties and catalyst
conversion of CO2 to methane and ethylene on a Cu-porphyrin electro­
microstructure of the cathode [353]. Design and control of surface fac­
catalyst in aqueous electrolyte [375].
eting and nanostructure of the catalytic cathode material profoundly
For CO2 mitigation purposes, it should be noted that the electro­
impact the selectivity and product distribution of CO2 reduction. These
chemical conversion of CO2 to useful chemicals and fuels becomes
challenges make it difficult to control product selectivity for the desired
meaningful and impactful only if renewable electricity is employed for
fuel. On the practical side, gas bubbling due to evolution reaction in
this process. Besides electrolytic conversion, other pathways involving
liquid electrolyte cells also poses efficiency and performance losses due
room temperature photocatalytic and photoelectrochemical reduction
to increased resistance of the electrolyte, blockage of catalytic surfaces
of CO2 to methanol and other hydrocarbons have also been reported in
by bubble formation and disruption to heat and mass transfer by rising
recent articles and reviews [376–380]. Obviously, there is clear
bubbles. More research is needed to address and overcome these tech­
advantage in utilizing solar energy for CO2 reduction and conversion.
nical barriers. Also, scaling up these processes to quantities meaningful
Most of these photon-based approaches employ aqueous electrolytes,
for CO2 mitigation measures is yet to be demonstrated. Developing these
but exhibit low conversion efficiency, current density, and selectivity
advanced technologies for cost effective conversion of CO2 to fuels still
values.
remains a desirable goal, albeit with major challenges.
CO2 electrolyzers also exhibit high polarization losses due to the
It is important to note that CO2 utilization may only complement
sluggish rates for the CO2 reduction and oxygen evolution reaction
mitigation at best. It cannot and should not be considered as a
(OER) at the cathode and anode, respectively. They usually operate
replacement or alternative to storage. Indeed, an earlier IPCC report
under high cell potentials in excess of 2 V with modest current densities

43
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

[312] had suggested that CO2 utilization may not make a significant price is low [384]. There are also geographic and structural differences
dent in reducing atmospheric emissions. Similarly, a more recent life and variations across the globe in implementing the carbon price. There
cycle analysis on the environmental impact of CCS and CCU reported are some regions like California in the U.S. as well as some nations,
that the global warming potential (GWP) of CCS was estimated at 276 especially in Europe, have been practicing cap-and trade for many years.
kgCO2eq/tCO2 removed, which was 215 times lower than the GWP of 59, There are few others like Northern European countries including Nor­
400 kgCO2eq/tCO2 for CCU for the production of a chemical such as way, Finland, and Denmark that implement a tax on carbon emissions.
dimethylcarbonate [94]. Nevertheless, integrating CO2 utilization to The variation of different forms of carbon pricing either implemented, in
renewable energy sources offers a promising pathway to utilize this planning or scheduled are mapped in Fig. 49, which also indicates that
valuable carbon source and close the carbon cycle for the production of large parts of the world have yet to take action.
useful materials, chemical feedstocks, fuels, and products. It should be noted that the global emissions covered under the car­
Incidentally, there is a useful resource to asses the feasibility and bon tax initiatives potentially amounted to 3 GtCO2e in 2020, which
environmental viability of CCU processes that may greatly benefit from corresponded to only 5.6% of global greenhouse gas (GHG) emissions,
life cycle analysis tools, namely, the NETL CO2U LCA Guidance Toolkit, while those covered under emission traded schemes (ETS) or cap-and-
recently developed and provided through the U.S. Department of En­ trade targeted 9 GtCO2e (or, 17% of global GHG), for a total of 12
ergy, National Energy Technology Laboratory (NETL) website [381]. GtCO2e or 22.3% of the global GHG emissions [385]. In reality, however,
The toolkit provides data, resources and modeling, as well as guidance only a total of 9 GtCO2e (or, 16% of global GHG) of the 12 GtCO2e has
and templates to funding recipients in performing life cycle analyses for actually been implemented in 2020. Nevertheless, there is agreement
evaluating the environmental impact of CO2 utilization ideas. that pricing carbon may work well for some sectors and industries with
large volume point source emissions such as electric power generation,
10. Pricing carbon fertilizer production and cement or steel manufacturing, while pricing is
difficult to implement for the transportation sector and agro-food
Mitigation measures incentivizing reductions in CO2 emissions also industries.
require placing a price on carbon. This can be an effective policy tool to On the other hand, the skeptics argue that carbon pricing has
promote investment and accelerate progress in decarbonizing the global inherent problems and most likely, its implementation may not be suf­
energy systems. In fact, many have viewed carbon pricing to be the most ficient to curb emissions and mitigate climate change [384]. It addresses
effective tool to curb CO2 emissions and also serve as the cornerstone for only a small fraction (~16%) of the global emissions. Assessing the
climate policies [382]. However, assessing and assigning a price on regional impact of the carbon price may also be complicated by external
carbon emissions is a complex problem that involves societal, economic, factors. For example, should one tax the emitting plant, or its corporate
policy, environmental and technical components. It is also important to owner located in another state or country. Furthermore, the price for
understand the social costs of carbon and how climate policies may carbon is naturally borne by all stakeholders in the chain including the
impact the economy [383]. business and industrial sectors as well as the communities and nations,
The “price” or “fee” placed on CO2 emissions also needs to be fair, albeit in varying degrees. It was argued that the cost of carbon pricing
transparent and uniform in its implementation. Otherwise there is a risk unfairly burdens lower income people. As a fraction of household in­
that some emitters may get a free ride on the sincere mitigation efforts of come, it was estimated that the lowest 10% income households in the U.
others, or heavy polluters may relocate to regions where the carbon S. would end up paying nearly 3 times more in carbon tax than the upper

Fig. 49. Map of regional and national distribution of carbon pricing initiatives at various stages of planning and implementation across the globe, including carbon
tax and emission traded schemes (ETS) i.e., cap-and-trade [385].

44
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 50. Implemented prices for carbon in various carbon pricing initiatives worldwide including carbon tax and emissions traded schemes (ETS), or cap-and-trade
[385]. Vertical axis is in units of US$/tCO2e.

10% of households in annual income [386]. This is because carbon processes such as iron and steel, cement, fermentation, etc. Some of
emitting activities and expenses such as heating and transportation them are large volume point emission sources, some are distributed, and
make up a proportionately larger fraction of the total income of poor yet some others, such as wildfires, preside over vastly large areas that
families than rich households. Hence, fairness is an important consid­ make it impossible to collect and capture CO2. Such diversity in the
eration. Moreover, the price of carbon should be sufficiently high to locations, modes, volumes, compositions, and origins of CO2 emissions
serve as an effective deterrent. Currently, it is difficult to claim that this demand equally diverse technologies, processes, methodologies and
is the case. Fig. 50 presents the widely varying price of carbon imple­ policies for CCS. There is no single technology that has the ability, ca­
mented in various regions and countries around the world in the form of pacity or maturity to address the multitudes of challenges posed by
tax or cap-and-trade schemes. It is clear that most of the prices lie below large-scale CCS from all these emission sources. Hence, a multi-prong
$25/tCO2e, and many are even considerably lower at around $10/tCO2 approach to CCS is essential for climate mitigation, in the same
to $15/tCO2e [385]. These numbers are significantly lower than the manner that we need a diverse portfolio of energy production and en­
actual cost of CCS as discussed in previous sections. So the current prices ergy storage technologies for both practical and energy security reasons
on carbon fail to provide a strong incentive and motivation to capture [101]. With impending and growing risks of climate change, it is ur­
CO2 emissions. This poses a serious problem for achieving climate gently imperative to develop a rich portfolio of CCS technologies to
mitigation. To meet the 2 ◦ C temperature rise target set by the 2015 address large-scale CO2 capture. Many of the major CCS technologies,
Paris Agreement, the 2017 Report of the High-Level Commission on processes, and materials have been discussed and reviewed in the pre­
Carbon Prices [387] recommended implementing a price of $40/tCO2 to sent article. With each passing year without any significant global action
$80/tCO2 by 2020, and $50/tCO2 to $100/tCO2 by 2030 to help close taken at-scale, the task of achieving the emission reduction goal of 25
the emissions gap. It was argued that voluntary measures for CCS will GtCO2e by 2030 set out by the UN Emissions Gap Report of 2019 [11]
not work but enforced policies will, and simply pricing carbon may also becomes more daunting and costly, and the required reduction rates
be insufficient to incentivize decarbonization of the energy sector and demand more steep cuts as depicted in Fig. 51. For example, if we had
deploy CCS at large scale, especially if the price of carbon is set too high started global emission cutting efforts back in 2010, we would have to
that penalizes the economy not just the fossil fuel-based electrical power reduce emissions by only 3.3 per year until 2030. If we started last year
sector [46]. The low price of carbon implemented in some form in select in 2020, it would have required annual reductions of 7.6% to achieve the
countries covers only a small fraction of global emissions, and this low 2030 emissions target of 25 GtCO2e. If we continue to procrastinate until
pricing is sometimes viewed as permission to pollute [388]. These 2025, the required annual reductions increase to 15.4%, which is a tall
considerations highlight the challenges for the global community and order for the world to meet. In other words, it will be significantly more
the importance of correct assessment for a fair, transparent and uniform expensive to wait.
price for carbon. Decarbonization of the energy economy and commensurate reduc­
tion of CO2 emissions clearly point to rapid transformation of the global
11. Assessments and outlook energy system to renewable energy sources, which have low carbon
intensities (or, CO2 footprint per unit of energy). Indeed, Table 18
Strong dependence of our energy systems on fossil fuels is the pri­ summarizes the carbon intensities of renewable, nuclear and fossil fuel-
mary culprit for CO2 emissions. Hence, massive adoption of renewable based power generation sources that were reported by IPCC [389] via
sources is critically important for climate mitigation, but it may not be life cycle analysis (LCA), indicating the significant advantage of
sufficient to achieve total decarbonization of global energy systems renewable power sources as well as nuclear when compared with fossil
[21]. The sources of CO2 emissions across the planet range from, to fuel-based power generation. Barring its nuclear waste problem, nuclear
name a few, forest and wildfires, to transportation vehicles, to building energy is the only carbon-free option to provide base-load electricity.
heating and cooling systems, to power generation plants, to industrial Intermittent renewable sources such as solar, wind, geothermal, and

45
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

Fig. 51. Required reductions in CO2 emissions recommended by the UN Emissions Gap Report 2019, with respect to the starting year of reductions in order to meet
the 25 GtCO2e by the year 2030 [11].

Table 18
Comparison of carbon intensities of power sources via life cycle analysis. Data
are compiled from Table A.II.4 of IPCC 2011, Special Report on renewable en­
ergy sources and climate change mitigation (SRREN) [391].
Power Technology Carbon Intensities of Power Generation Sources
(kgCO2-eq/MWhe)

25% to 75% percentile range 50% percentile

Coal 877 to 1130 1001


Natural Gas 422 to 548 469
Biopower 360 to 37 18
Nuclear 8 to 45 16
Hydropower 3 to 7 4
Wind 8 to 20 12
Solar PV 29 to 80 46
Solar Thermal 14 to 32 22
Geothermal 20- 57 45

marine are not stand-alone, dispatchable energy sources but instead


require integration with adequate energy storage capacities. Li-ion
rechargeable batteries emerged to become the preferred electricity
storage systems widely employed today. However, life cycle analysis of
Li-ion battery-based storage indicates that 1 kWh of storage capacity
requires an energy expenditure of 400 kWh over the 20-year life time of
battery storage, resulting in emissions of 75 kgCO2 (or 75,000
kgCO2/MWh) [390], while a similar LCA study estimated the carbon
Fig. 52. Collective carbon intensities between 2010 and 2050 of OECD (lower
intensity of 1 kWh of Li-ion battery storage to be 110 kgCO2 (or 110,000 curve) and non-OECD (upper curve) nations [26]. (OECD: Organization for
kgCO2/MWh)[393]. The results suggest that renewable storage is nearly Economic Co-operation and Development). Also note: 1 Btu = 1055.06 J.
100 times higher in carbon intensity than fossil fuel-based power gen­
eration, which negates the advantage of renewable energy presented in
capacities ranging over several decades of time scales from sub-minutes
Table 18.
to hundreds of hours must be developed aggressively to overcome the
When assessing the impact of energy technology options, it is
inherently intermittent, variable, unpredictable and nondispatchable
therefore important to consider the energy penalty and CO2 emissions
nature of most renewable energy sources [101,392,394].
along the entire value chain of energy systems, devices and technologies
There is also uncertainty if 100% renewables-based energy economy
in choosing the most environmentally friendly, cost effective and effi­
is achievable, or even realistic. Claims of total decarbonization of the
cient option. In that regard, net energy analysis may be an effective tool
global energy economy by banking on 100% renewable sources [18,19]
to account for carbon, assess sustainability and verify the impact of
were met with skepticism and viewed not to be realistic for complete
energy and climate policies on the society, environment, economy and
elimination of CO2 emissions [20,21]. Although efficiency improvement
climate [392]. For example, it was estimated that the return of energy
measures and moving away from coal-fired power generation are ex­
investment in wind power would return 80 times the energy consumed
pected to help lower the global carbon (and also, energy) intensity (see
for manufacturing over its lifetime, while solar photovoltaic power
Fig. 52), world’s energy-related CO2 emission is still expected to grow at
returns only 10 times due to its significantly higher manufacturing cost
an annual rate of 0.6% through 2050 [26]. Hence, rapid development
[393]. Regardless, however, utility-scale deployment of renewable
and deployment of cost effective and energy efficient CCS technologies
sources to completely displace and replace fossil fuel-based energy and
still presents an urgent need to curb atmospheric CO2 levels and mitigate
power systems is in part constrained by the current state of the global
climate change for a sustainable future.
electrical grid, but it is also constrained by the availability of cost
Some of the primary components along the CCS process chain,
effective, scalable, efficient, and reliable energy storage options. In other
including CO2 separation and capture, compression for transport, and
words, utility scale energy storage technologies with discharge

46
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

finally storage, are commercially available, and some are still under based solvent scrubbing has gained the most interest in practice,
development with varying energy penalties and cost estimates that are partly because it is a commercial technology that is already employed
site or process specific. Although there is sufficient industrial experience for gas cleanup in industrial applications. Similar to other options,
with many of the individual components in the process chain, for any however, amine-based CO2 capture is also an energy intensive and
CCS pathway to be economically viable and widely deployed at a scale expensive process. The minimum theoretical work required for CO2
commensurate with the volume of CO2 emissions, systems level inte­ capture using amine solvents followed by compression to 15 MPa is
gration of its components is essential. Unfortunately, industrial experi­ estimated to be around 420 MJ/tCO2, where more than half of this en­
ence with integration of the CCS chain is lacking at the gigaton scale. ergy penalty is due to the CO2 recovery step during solvent regeneration,
In its 2014 report, the United Nations Environment Programme and the remaining is consumed during the compression process [216].
(UNEP) had considered the most cost-effective scenarios to curb CO2 Effective integration of thermal processes, such as solvent regeneration
emissions based on several assumptions, namely, 1) all countries begin during post-combustion CO2 capture with fossil fuel-based power plants
mitigation measures immediately, 2) there is a single global price on is an important consideration to improve overall efficiency, but this is
carbon, and 3) all key technologies are available [395]. The UNEP report also impacted adversely by the humidity content of the inlet flue stream
indicated that limiting the greenhouse gas concentrations in the atmo­ [80]. Hence, dehumidifying the flue gas stream before the CO2 capture
sphere to around 450 ppmv CO2e clearly quantifies the momentous step helps lower the energy burden, contactor area, and capital cost.
challenge of reducing emissions by a factor of nearly two in order to Although there is an energy penalty of 8% to 12% for oxy-combustion
have a chance to stay within the 1.5 to 2.0 ◦ C temperature rise above the due primarily to separating oxygen from air [113,131], this technology
pre-industrial levels. Similarly, IPCC studied the technological options offers not only reduced SOx and NOx emissions and ease of capture for
to limit atmospheric CO2 concentrations not to exceed 450 to 530 ppmv CO2 but also higher plant efficiency and a lower threshold for retrofit­
by 2100 [396]. Many different technologies and their CO2 emission ting oxy-combustion into existing fossil fuel-based power plants.
intensities have been analyzed for their impact on atmospheric CO2 Direct air capture, certainly a costly and highly energy intensive
concentrations as well as the resulting cost of electricity, which increases option, is considered not only to complement CCS policies but also as a
with implementation of CCS as expected. Their 2014 analyses predicted potential countermeasure to capture CO2 emissions from uncontrolled
that utility scale solar PV and wind would compete favorably with fossil events such as coal mine fires, forest fires, and wildfires, or distributed
fuel-based power, which recently became true indeed [397]. Even sources such as transportation vehicles, or leakages from CO2 storage
though fuel cell power generation is inherently highly efficient and of­ sites. Arguably, direct air capture presents technically the most chal­
fers competitive cost of electricity (see Fig. 24 above), the IPCC 2014 lenging option in mitigating CO2 emissions. Although there is com­
report did not give any consideration to this promising technology op­ mercial interest in this nascent approach, the future prospects of
tion for power production. More importantly, CO2 can readily be developing and adopting DAC widely at the gigaton scale remains un­
captured from fuel cell power plants without post separation operations certain. To achieve negative emissions, its success may rely on the use of
and eliminate the need for parasitic power demand for separation. non-carbon energy sources such as renewables and nuclear to meet the
In this regard, this author believes it is imperative to initiate efforts high energy demand for DAC.
to transition the conventional fossil fuel-based power generation tech­ Although CO2 utilization does not seem to have the magnitude of
nologies from air-based to oxygen-based advanced, efficient power scale in gigatons, it nevertheless provides a valuable carbon source for
generation that also offer capture-ready production of CO2 flue streams, the synthesis and manufacturing of a wide range of materials, chemical,
as discussed above in this article [22]. All credible projections provided fuels and products routinely needed and utilized by the global com­
in this article clearly indicate that fossil fuels are expected to play a munity. Using CO2 as the C1 building block also helps close the carbon
major role in the global energy mix well into the foreseeable future. cycle by recycling it back into useful products instead of deriving the
Although their share in electricity generation will decrease, the amount required carbon from fossil fuel sources.
of fossil fuel use actually will continue to increase (see Figs. 4 and 5). The low hanging fruit of increasing the efficiency of energy pro­
China (~28% of global emissions) has already surpassed the U.S. (~15% duction, conversion and storage processes is also a critically important
of global emissions) as the largest CO2 emitter, and while investing in consideration for CO2 mitigation. Advanced power generation technol­
renewables also makes large investments in fossil fuel based power ogies such as renewables and fuel cells to increase both the production
generation. China also is investing heavily in its Belt and Road Initiative and end-use efficiency of electricity generation provides additional op­
that involves building nearly 140 coal-fired power plants across Asia and tions to compensate for the energy penalty paid during carbon capture
Eurasia. For all the reasons discussed and reviewed in the present article, and storage process. It was estimated that increasing the electrical ef­
it is too risky to stick with conventional air-based combustion technol­ ficiency by a modest 15% to 20% would suffice to make up the power
ogies for power generation as if that is the permanently ingrained law of that would be expanded for CCS [106].
the land. This needs to change. Implementing a global, fair and uniform As recommended by the 2019 UN Emission Gap Report [11] for full
price for carbon emissions will help incentivize private investments and decarbonization especially of the energy and transportation sectors,
federal support in developing, adapting and transitioning to rapid electrification across various energy sectors while phasing out
oxygen-based power generation technologies in a parallel effort that fossil fuels is another viable strategy to curb CO2 emissions. This in­
complements global mitigation measures. volves shifting most industrial processes towards electricity, decarbon­
By its nature, CCS is a multi-faceted, complex and arguably an izing building heating and cooling systems via electrification and
existential technology that is intimately interlinked with and conse­ efficiency enhancements, agricultural measures to shift from
quential to almost every human activity on this planet. Regardless of the animal-based to plant-based food sources, net-zero deforestation pol­
choice of particular CCS technology, materials or systems under icies to protect natural ecosystems, and restoration and reforestation
consideration, all have their advantages and shortcomings, and all measures to improve land carbon stocks for increased CO2 uptake.
require vast amounts of energy for CCS. Equally important is the fact It is likely that technical efforts towards full decarbonization of the
that all options suffer also from knowledge gaps and experience espe­ energy economy by the end of this century may fall short of addressing
cially in scaling up at least some of these CCS options to multi-gigaton the atmospheric CO2 problem alone. Given the largest carbon flux be­
CO2 per year capacity (see Table 3). tween land and the atmosphere of nearly 440 GtCO2/yr occurs via
As discussed extensively in this article, major barriers to CCS are the photosynthesis [73], large scale exploitation and management of the
cost and high energy penalty of capturing CO2 from flue streams and to terrestrial biosphere and the biological carbon cycle including affores­
compress it further for transportation and storage. Among the many tation and reforestation, land and marine biota, as well as mineralization
technological options for CO2 capture discussed in this article, amine- and rock weathering, animal husbandry, food supply chains are fertile

47
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

areas to make large-scale impact on CO2 mitigation strategies. are burdened with technical, environmental, financial, policy, and social
Even though fair assessment of carbon pricing is challenging, it challenges. This high degree of complexity and uncertainty naturally
nevertheless conveys and represents the cumulative external costs of increases investment risks and also hinders development efforts. For
climate mitigation for all stakeholders. It also educates the public, in­ sure, the road to commercial CCS technologies is riddled with difficult
dustry and policy makers and brings to fore the urgency of addressing challenges and choices. Furthermore, lack of a globally agreed uniform
imminent danger of catastrophic climate change, and implement mea­ pricing structure on carbon makes any commercial undertaking politi­
sures to integrate decarbonization strategies and programs into action­ cally and financially risky. The enormous magnitude, complexity, and
able policies. So carbon pricing should be considered as part of the the time line of the problem makes it difficult for private sector to invest
overall mix of measures and policies that promote reduction in CO2 heavily in emerging technologies that complement renewables,
emissions and innovation carbon capture technologies. including fuel cells for efficient power generation from fossil fuels with
Most of the technologies and processes discussed in the present capture-ready CO2 production. Such economic and technological risks
article are not yet ready for commercial implementation, with the are certainly higher than private investments can bear. So federal pro­
exception of several including oxy-combustion, IGCC and amine-based grams, incentives, and assistance across the globe are needed to lower
solvent scrubbing of CO2. Even then, several amine scrubbing, IGCC, the hurdles and gain better insight in overcoming the technical and cost
and oxy-combustion projects have failed to demonstrate economic barriers for CCS.
viability [158,160]. Some have abandoned their carbon capture activity
[398]. The barriers to commercialization are complex but mostly 12. Key challenges and opportunities
economical, technical and political in nature. Cost is still a big issue, and
unless major incentives are provided, private industry will shy away CCS clearly faces a diverse range of challenges on all fronts, ranging
from shouldering the risks alone. Technical barriers are multiple and from policy and regulatory issues, to financial and economic, to climate
each particular technology poses many challenges yet to be resolved. and environment, to quality of life and health, to social and income
Political barriers across the globe failed to forge a unified and urgent inequalities, and global agreement on a fair and sound strategy to deal
response to mitigate climate change. Unfortunately, these issues paint a with mitigation. On the technical side, the diversity of technologies,
rather bleak outcome, and in all likelihood, the world may not be able to processes, and materials along the CCS and CCU chains present addi­
comply with the IPCC guidelines [399] to cut CO2 emissions by 55% by tional hurdles, knowledge gap and challenges that require fundamental
the year 2030 in order to limit global warming to within 1.5 ◦ C [12,13]. research and understanding as well as engineering solutions to over­
Unfortunately, economic arguments have often dominated and come them. Although many of these issues are addressed and discussed
diverted the public discourse away from the ultimate goal of avoiding under their respective sections, some of the more visible challenges are
imminent risks to life on planet Earth. In general, the global public grouped and summarized below in random order:
opinion is far ahead of the world’s governments, who in large part have
failed to demonstrate the political will and the commitment to enact - Advances in functional materials that offer superior properties
investments commensurate with the massive volumes required for CCS. ranging from more efficient and cost-effective separation of oxygen
It is likely an even bigger price will have to be paid later if the world from air and CO2 from flue gas streams, to novel liquid and solid
continues to delay aggressive action and collective implementation of sorbents that offer higher CO2 uptake capacity with less environ­
effective CCS policies to limit the temperature rise to within 1.5 ◦ C. mental impact are needed to expand the technology options for CCS.
Indeed, it is more expensive to do nothing now. Impact of climate - Fundamental understanding of CO2-sorbent interactions and control
change on food production and supplies, displacement of large pop­ of surface binding processes that govern preferential uptake of CO2
ulations due to sustained draft and water shortages, flooding and loss of from flue gasses will expedite progress in designing advanced ma­
land along low lying coastal regions by rising sea levels are plausible terials with superior sorption properties.
scenarios of existential consequences. - Theoretical simulations will advance fundamental understanding of
To illustrate the immense price of little or no action now but pay binding mechanisms at the atomic scale as well as expedite materials
more later, consider the impact of climate change on global commu­ screening and selection processes.
nities in both direct and hidden costs. Between 1980 and 2018, the U.S. - Similarly, implementation of advanced materials processing tools
has suffered 241 climate-related natural disasters that caused more than and techniques to tune, modify and alter materials properties by
$1 billion in damages per incident, for a total loss of 13,188 lives and nanostructuring and control of microstructure will help improve
cumulative cost of more than $1.6 trillion, or more than $42 billion per desired properties critical for removing CO2 from gaseous streams.
year [400]. More recently, the average price tag for natural disasters in - Development of non-aqueous solvents with higher CO2 affinity will
the U.S. totaled to more than $99 billion annually between 2014 and help improve uptake capacity.
2018. By contrast, the most recent fiscal year 2021 (FY2021) allocation - Development of scalable, stable and environmentally benign sol­
for basic energy research funding for the U.S. DOE Office of Science, vents, sorbents and processes for CO2 capture will help lower the
Basic Energy Sciences was $2.25 billion [401]. Similarly, the cumulative threshold for social and political barriers for global implementation
totals for all energy related federal funding in the U.S. has been of CCS at large scale.
reasonably steady around $3 billion to $4 billion annually, which is - Pre-combustion and post-combustion capture both suffer from heavy
nearly 3% of the annual damages caused by natural disasters. This stark energy penalties.
comparison and large discrepancy represents a ‘pay now or pay more - Direct air capture (DAC) demands the highest energy penalty, and is
later’ choice. Obviously, basic research is not capable of solving all arguably the most costly, but still deserves attention as it removes
problems natural or man-made, but it will guide and provide techno­ CO2 from the atmosphere regardless of the origin, nature or geog­
logical options to solve pressing societal and global problems. Given the raphy of the emission source. Advances in materials and solvents
intimate interaction and interdependence of energy and climate change made for post-combustion capture will also benefit DAC greatly.
(see Fig. 2), this raises the question if enough investments are being - Improving efficiencies of individual steps along the combustion-
wisely made by policy makers to support development of advanced power generation-CCS process chain will help reduce emissions as
technologies to address energy and climate change issues collectively well as the financial, technical and energy burden of CO2 capture.
and strategically. - Coupling renewable energy sources and fuel cells into the power
As pointed out in a recent U.S. National Academies workshop generation-CCS process chain will also help lower the energy penalty
regarding deep decarbonization of economies [402], technology options of CCS.
for rapid commercialization and deployment of CO2 capture and storage

48
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

- Developing scalable CCS processes, materials and technologies will interests or personal relationships that could have appeared to influence
benefit from economies of scale and help lower costs. the work reported in this paper.
- Better understanding of the geochemistry of CO2-soil interactions
will help more accurate assessment of long-term storage sites. Acknowledgements
- Developing earth abundant catalysts and understanding of elemen­
tary mechanisms of CO2 reduction on catalytic surfaces to higher The author gratefully acknowledges partial support from the Stan­
hydrocarbons and fuels will enable utilization and recycling of this ford Energy 3.0 program, and helpful discussions with Prof. F. B. Prinz of
valuable carbon source to make useful products. Stanford University.
- The magnitude of CCS is enormous in terms of volumes required in
the tens of gigatons scale. This is at a scale unlike the industrial ca­ References
pacities ever built before. This endeavor will require a different en­
gineering approach to meet the unprecedented magnitude of the [1] The Washington Post, Dec. 5, 2018. https://www.washingtonpost.com/ener
problem. gy-environment/2018/12/05/we-are-trouble-global-carbon-emissions-reache
d-new-record-high/ (accessed Nov. 27, 2020).
- CCS is an expensive proposition and currently the cost for CCS is [2] United Nations Climate Change Conference, COP24-KATOWICE 2018, http://cop
generally much higher than most of the cap-and-trade prices or taxes 24.gov.pl.
on carbon. Even with regulation, this imbalance encourages in­ [3] The Guardian, Dec. 2, 2020, https://www.theguardian.com/environment/2020/
dec/02/humanity-is-waging-war-on-nature-says-un-secretary-general-antonio-
dustries to elect the lower cost to pollute and emit, and does not guterres.
provide the necessary incentive for industries, governments and the [4] The Production Gap Report 2020, by UN Environment Programme (UNEP),
private sector to invest in developing technological solutions. This Stockholm Environment Institute (SEI), International Institute for Sustainable
Development (IISD), Overseas Development Institute (ODI), and E3G, http://
loophole must be plugged. productiongap.org/2020report/.
- Pricing carbon emissions is certainly an effective deterrent, but [5] UN Emissions Gap Report 2019 (November 2019), by UN Environment
placing a fair, equitable, and uniform price agreeable to citizens and Programme (UNEP), https://wedocs.unep.org/bitstream/handle/20.500.11822/
30797/EGR2019.pdf?sequence=1&isAllowed=y (accessed June 15, 2021).
governments of the global community is also a complex and politi­
[6] Global CO2 emissions in 2019 (Feb. 11, 2020), International Energy Agency,
cally complicated problem with many potholes. https://www.iea.org/articles/global-co2-emissions-in-2019 (accessed Feb. 25,
- Climate change is arguably an existential challenge and inaction is 2020).
[7] BP Statistical Review of World Energy 2019 (68th edition), https://www.bp.
the most expensive choice we can make.
com/content/dam/bp/business-sites/en/global/corporate/pdfs/energy-econom
ics/statistical-review/bp-stats-review-2019-coal.pdf (accessed Feb 25, 2020).
13. Summary [8] British Petroleum, https://www.bp.com/content/dam/bp/business-sites/en/g
lobal/corporate/pdfs/energy-economics/statistical-review/bp-stats-review-201
9-co2-emissions.pdf (accessed Feb 25, 2020).
This overview covers and reviews a broad range of technologies, [9] U.S. Environmental Protection Agency, Global Green House Gas Emissions Data,
processes and materials for carbon capture and storage (CCS) as well as https://www.epa.gov/ghgemissions/global-greenhouse-gas-emissions-data
utilization, and critically examines their technological, economic, and (accessed June 12, 2021).
[10] IPCC-SR15 Technical Summary, https://www.ipcc.ch/sr15/technical
environmental merits. The urgency and consequential importance of -summary/(accessed May 24, 2021).
CO2 mitigation is stressed. Wide-ranging technologies towards decar­ [11] UN Emissions Gap Report 2019 (November 2019), by UN Environment
bonization of fossil fuel-based power generation with full carbon capture Programme (UNEP), https://wedocs.unep.org/bitstream/handle/20.500.11822
/30797/EGR2019.pdf (accessed Dec 5, 2020).
are discussed. Mature as well as emerging technological options both for [12] IPCC-SR15 Technical Summary, www.ipcc.ch/pdf/special-reports/sr15/sr15_ts.
CO2 capture and sequestration towards long term storage, as well as pdf.
utilization of CO2 are assessed. Although there seems to be more than [13] IPCC Special Report 2018- Global Warming of 1.5 ◦ C, https://www.ipcc.ch/sr15/
(accessed May 25, 2021).
sufficient capacity for CO2 storage around the world, persistent political [14] International Energy Outlook 2020 (IEO2020), U.S. Energy Information
and economic hurdles, knowledge gaps, and lack of industrial experi­ Administration, https://www.eia.gov/outlooks/ieo/ (accessed March 25, 2021).
ence in large scale and long term CO2 storage and sequestration in [15] Gür TM. Review of electrical energy storage technologies, materials, and systems:
challenges and prospects for large-scale grid storage. Energy Environ. Sci. 2018;
geological structures hinder rapid progress and implementation.
11:2696–767.
As expected, all options for CCS are energy intensive and carry a [16] Energy Storage Grand Challenge: Energy Storage Market Report. NREL/TP-5400-
price tag. There is no silver bullet, no single technology to resolve the 78461, DOE/GO-102020-5497. National Renewable Energy Laboratory, U.S.
climate mitigation problem. The quantities involved in CCS are also Department of Energy; Dec. 2020. https://www.energy.gov/sites/prod/
files/2020/12/f81/Energy%20Storage%20Market%20Report%202020_0.pdf.
massive and far exceed our current industrial capacities. The diversity of accessed May 2, 2021.
the compositions, volumes, locations and types of CO2 emissions also [17] International Renewable Energy Agency (IRENA), https://www.irena.org/Statisti
preclude singular solutions. Hence a multi-prong strategy is required to cs/View-Data-by-Topic/Capacity-and-Generation/Statistics-Time-Series.
[18] Jacobson MZ, Delucchi MA, MA Cameron, Frew BA. Low-cost solution to the grid
rapidly address the impending risk of irreversible climate change. reliability problem with 100% penetration of intermittent wind, water, and solar
Multiple technology options need to be explored, developed at the for all purposes. Proceed. National Acad. Sci. 2015;112:15060–5.
gigaton-scale and rapidly deployed to address this arguably existential [19] Jacobson MZ, Delucchi MA, Bauer ZAF, Goodman SC, Chapman WE,
Cameron MA, Bozonnat C, Chobadi L, Clonts HA, Enevoldsen P, Erwin JR,
challenge to the wellbeing of the ecosystem of our planet. As this multi- Fobi SN, Goldstrom OK, Hennesy EM, Liu J, Lo J, Meyer CB, Morris SB, Moy KR,
dimensional problem is a highly risky endeavor for the private sector to O’Neill PN, Petkov I, Redfern S, Schucker R, Sontag MA, Wang J, Weiner E,
take on alone, federal governments around the globe need to step in and Yachanin AS. 100% clean and renewable wind, water, and sunlight all-sector
energy roadmaps for 139 countries of the world. Joule 2017;1:108–21.
lower the risk barriers.
[20] Clack CTM, Qvist SA, Apt J, Bazilian M, Brandt AR, Cladeira K, Davis SJ,
Among all technological options, amine-based chemical absorption Diakov V, Handschy MA, Hines PDH, Jaramillo P, Kammen DM, Long JCS,
is commercially more mature and provide comparatively a lower com­ Morgan MG, Reed A, Sivaram V, Sweeney J, Tynan GR, Victor DG, Weyant JP,
Whitacre JF. Evaluation of a proposal for reliable low-cost grid power with 100%
mercial barrier for CO2 capture, albeit not without its own economic,
wind, water, and solar. Proceed. National Acad. Sci. 2017;114:6722–7.
environmental, and technical challenges. On the other hand, oxygen- [21] de Chalendar JA, Benson SM. Why 100% renewable energy is not enough. Joule
based combustion, pre-combustion and fuel cell-based energy conver­ 2019;3:1389–93.
sion processes discussed in this article offer alternative solutions by [22] Gür TM. Perspectives on oxygen-based coal conversion towards zero-carbon
power generation. Energy 2020;196(117074):1–9.
eliminating the need for CO2 separation from flue gas streams. [23] Gür TM. Progress in carbon fuel cells for clean coal technology pipeline. Intern. J.
Energy Res. 2016;40:13–29.
Declaration of Competing Interest [24] The Water-Energy Nexus: Challenges and Opportunities (June 2014), U.S.
Department of Energy, https://www.energy.gov/sites/prod/files/2014/07/f17
/Water%20Energy%20Nexus%20Full%20Report%20July%202014.pdf (accessed
The author declares that he has no known competing financial June 22, 2021).

49
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

[25] We’re at a turning point on climate change. But most countries are still choosing [52] IPCC Special Report on Carbon Dioxide Capture and Storage, Chapter 2. Sources
fossil fuels over clean energy, report says, CNN News, Dec. 2, 2020, https://www. of CO2, https://www.ipcc.ch/site/assets/uploads/2018/03/srccs_chapter2-1.pdf
cnn.com/2020/12/02/world/climate-production-gap-fossil-fuels-intl-hnk/index. (accessed May 26, 2021).
html. [53] Greenhouse gas emissions from a typical passenger vehicle. U.S. Environmental
[26] International Energy Outlook 2019, US Energy Information Administration, Sept Protection Agency; 2021. https://nepis.epa.gov/Exe/ZyPDF.cgi?Dockey=P100
24, 2019, https://www.eia.gov/outlooks/ieo/pdf/ieo2019.pdf (accessed March U8YT.pdf. accessed June 5.
4, 2020). [54] Carbon budget and trends 2018, Global Carbon Project report (Dec. 5, 2018),
[27] CarbonBrief. https://www.carbonbrief.org/mapped-worlds-coal-powerplants. www.globalcarbonproject.org/carbonbudget, (accessed May 19, 2021).
[28] The National Interest (Aug. 18, 2019), https://nationalinterest.org/feature/ch [55] Global Energy and CO2 Status Report 2017, International Energy Agency, report
inas-belt-androad-plan-destroying-world-74166. # OECD/IEA 2018 (March 2018).
[29] U.S. Energy Information Administration. Average tested heat rates by prime [56] Figueres C, Le Quere C, Mahindra A, Bate O, Whiteman G, Peters G, Guan D.
mover and energy source 2007–2012, http://www.eia.gov/electricity/annual/h Emissions are still rising: ramp up the cuts. Nature 2018;564:27–30.
tml/epa_08_02.html. [57] Friedlingstein P, O’Sullivan M, Jones MW, Andrew RM, Hauck J, Olsen A,
[30] Gür TM. Critical review of carbon conversion in “Carbon Fuel Cells. Chem. Rev. Peters GP, Peters W, Pongratz J, Sitch S, Le Quere C, Canadell JG, Ciais P,
2013;113:6179–206. Jackson RB, Alin S, Aragao LEOC, Almut A, Arora V, Bates NR, Becker M, Benoit-
[31] U.S. Department of Energy, Secretary of Energy Advisory Board (SEAB) Task Cattin A, Bittig HC, Bopp L, Bultan S, Chandra N, Chevallier F, Chini LP, Evans W,
Force report, CO2 utilization and negative emissions technologies, Dec. 13 Florentie L, Forster PM, Gasser T, Gehlen M, Gilfillan D, Gkritzalis T, Gregor L,
(2016), https://energy.gov/sites/prod/files/2016/12/f34/SEAB-CO2-TaskForce Gruber N, Harris I, Hartung K, Haverd V, Houghton RA, Ilyina T, Jain AK,
-FINAL-with%20transmittal%20ltr.pdf (accessed May 7, 2020). Joetzjer E, Kadono K, Kato E, Kitidis V, Korsbakken JI, Landschutzer P, Leferve N,
[32] Climate intervention: carbon dioxide removal and reliable sequestration. Wash., Lenton A, Lienert S, Liu Z, Lombardozzi D, Marland G, Metzl N, Munro DR,
D.C.: National Research Council, The National Academies Press; 2015. Nabel JEMS, Nakaoka S-I, Niwa Y, O’Brien K, Ono T, Palmer PI, Pierrot D,
[33] Bui M, Adjiman CS, Bardow A, Anthony EJ, Boston A, Brown S, Fennel PS, Fuss S, Poulter B, Resplandy L, Robertson E, Rodenbeck C, Schwinger J, Seferian R,
Galindo A, Hackett LA, Hallett JP, Herzog HJ, Jackson G, Kemper J, Krevor S, Skjelvan I, Smith AJP, Sutton AJ, Tanhua T, Tans PP, Tian H, Tilbrook B, van der
Maitland GC, Matuszewski M, Metcalfe IS, Petit C, Puxty SA, Reimer J, Willis G, Vuichard N, Walker AP, Wanninkhof R, Watson AJ, Willis D,
Reiner DM, Rubin ES, Scott SA, Shah N, Smit B, Martin Trusler JP, Weble P, Wiltshire AJ, Yuan W, Yue X, Zaehle S. Global carbon budget 2020. Earth Syst.
Wilcox J, Mac Dowell N. Carbon capture and storage (CCS): the way forward. Sci. Data 2020;12:3269–340. https://doi.org/10.5194/essd-12-3269-2020. 32-
Energy Environ. Sci. 2018;11:1062–176. 69-3340, see also.
[34] Davis SJ, Lewis NS, Shaner M, Aggarwal S, Arent D, Azevedo IL, Benson SM, [58] Global Energy Review. CO2 emissions. International Energy Agency; 2021.
Bradley T, Brouwer J, Chiang Y-M, Clark CTM, Cohen A, Doig S, Edmonds J, https://www.iea.org/reports/global-energy-review-2021/co2-emissions.
Fennell P, Field CB, Hannegan B, Hodge B-M, Heffert MI, Ingersoll E, Jaramillo P, accessed May 25, 2021.
Lackner KS, Mach KJ, Mastrandrea M, Ogden J, Peterson PF, Sanchaz DL, [59] Carbon dioxide levels hit new record; covid impact ‘a tiny blip’, wmo says. UN
Sperling D, Stagner J, Trancik JE, Yang C-J, Caldeira K. Net-zero emissions energy News; Nov. 23, 2020. https://news.un.org/en/story/2020/11/1078322.
systems. Science 2018;360. eaas9793 1-9. [60] The Production Gap Report 2020, UN Environment Programme, by UN
[35] Song C, Liu Q, Deng S, Zhao J, Li Y, Song Y, Li H. Alternative pathways for Environment Programme (UNEP), Stockholm Environment Institute (SEI),
efficient CO2 capture by hybrid processes – A review. Renew. Sustain. Energy International Institute for Sustainable Development (IISD), Overseas
Rev. 2018;82:215–31. Development Institute (ODI), and E3G, https://www.unenvironment.org/resourc
[36] Koytsoumpa EI, Bergins C, Kakaras E. The CO2 economy: review of CO2 capture es/report/production-gap-2020 (accessed Dec. 6, 2020).
and reuse technologies. J. Supercritical Fluids 2018;132:3–16. [61] Global Monitoring Laboratory, Earth Systems Research Laboratories, U.S.
[37] Kenarsari SD, Yang D, Jiang G, Zhang S, Wang J, Russell AG, Wei Q, Fan M. Department of Commerce, National Oceanic & Atmospheric Administration, https
Review of recent advances in carbon dioxide separation and capture. RSC Adv ://gml.noaa.gov/ccgg/trends/mlo.html (accessed May 24, 2021).
2013;3:22739–73. [62] Global Monitoring Laboratory, Earth Systems Research Laboratories, U.S.
[38] Garcia-Garcia G, Fernandez MC, Armstrong K, Woolass S, Styring P. Analytical Department of Commerce, National Oceanic & Atmospheric Administration,
review of life-cycle environmental impacts of carbon capture and utilization https://gml.noaa.gov/ccgg/trends/gr.html (accessed May 24, 2021).
technologies. ChemSusChem 2021;14:995–1015. [63] IPCC-AR5 report 2014 (Mitigation of Climate Change - 5th Assessment Report):
[39] Lee S-Y, Park S-J. A review on solid adsorbents for carbon capture. J. Indust. Eng. https://www.ipcc.ch/pdf/assessment-report/ar5/wg3/ipcc_wg3_ar5_technical-su
Chem. 2015;23:1–11. mmary.pdf.
[40] Rubin ES, Mantripragada H, Marks A, Versteeg P, Kitchin J. The outlook for [64] IPCC. Summary for Policymakers. Global warming of 1.5 ◦ C. an ipcc special
improved carbon capture technology. Progress Energy Combust. Sci. 2012;38: report on the impacts of global warming of 1.5 ◦ C above pre-industrial levels and
630–71. related global greenhouse gas emission pathways, in the context of strengthening
[41] Leeson D, Mac Dowell N, Shah N, Petit C, Fennell PS. A techno-economic analysis the global response to the threat of climate change, sustainable development, and
and systemic review of carbon capture and storage (CCS) applied to the iron and efforts to eradicate poverty. Geneva, Switzerland: World Meteorological
steel, cement, oil refining, and pulp and paper industries, as well as other high Organization; 2018. p. 32. Masson-Delmotte, V, Zhai P, Pörtner H-O, Roberts D,
purity sources. Intern. J. Greenhouse Gas Control 2017;61:71–84. Skea J, Shukla PR, Pirani A, Moufouma-Okia W, Péan C, Pidcock R, Connors S,
[42] Global Greenhouse Gas Emissions Data, U.S. Environmental Protection Agency, Matthews JBR, Chen Y, Zhou X, Gomis MI, Lonnoy E, Maycock T, Tignor M,
https://www.epa.gov/ghgemissions/global-greenhouse-gas-emissions-data Waterfield Talso available at, https://www.ipcc.ch/sr15/chapter/spm/. accessed
(accessed June 12, 2021). March 4, 2020.
[43] IPCC 2007, Contribution of Working Group I to the Fourth Assessment Report, [65] Irfan U, Zarracina J. Antarctica has lost 2.71 trillion tons of ice. Here is what that
“Climate Change 2007: the Physical Science Basis”, https://www.ipcc.ch/report looks like. Vox June 28, 2018. https://www.vox.com/science-and-health/2018/
/ar4/wg1/. 6/28/17475342/Antarctica-ice-melt-thaw-climate-clonge-sea-level.
[44] IPCC. In: Parry ML, Canziani OF, Palutikof JP, van der Linden PJ, Hanson CE, [66] Trusel LD, Das SB, Osman MB, Evans MJ, Smith BE, Fettweis X, McConnell JR,
editors. Contribution of working group ii to the fourth assessment report of the Noel BPY, van den Broeke MR. Nonlinear rise in Greenland runoff in response to
intergovernmental panel on climate change. Cambridge, United Kingdom, and post-industrial Arctic warming. Nature 2018;564:104–8.
New York, NY, USA: Cambridge University Press; 2007. http://www.ipcc.ch/ [67] Mann ME. Earth will cross the climate danger threshold by 2036. Sci. Am. 2014;
publications_and_data/ar4/wg2/en/contents.html. 310(4). 18 March 2014, https://www.scientificamerican.com/article/earth
[45] Socolow R., Desmond M., Aines R., Blackstock J., Bolland O., Kaarsberg T., Lewis -will-cross-the-climate-danger-threshold-by-2036/. accessed June 5, 2021.
N., Mazzotti M., Pfeffer A., Sawyer K., Siirola J., Smit B., Wilcox J. (2011), Direct [68] The developing world has hit the brakes on clean energy, MIT Technology
air capture of CO2 with chemicals: a technology assessment for the APS panel on Review, Nov. 25 (2019), see https://www.technologyreview.com/2019/11/25/
public affairs, American Physical Society Report (June 1, 2011). http://www.aps. 131845/the-developing-world-has-hit-the-brakes-on-clean-energy/.
org/policy/reports/assessments/upload/dac2011.pdf (accessed May 7, 2020). [69] Caldeira K, Davis SJ. Accounting for carbon dioxide emissions: a matter of time.
[46] Haszeldine RS. Carbon capture and storage: how green can black be? Science Proceedings National Academy Sciences 2011;108:8533–4.
2009;325:1647–52. [70] Davis SJ, Peters GP, Cladeira K. The supply chain of CO2 emissions. Proceedings
[47] Understanding Global Warming Potentials, U.S. Environmental Protection National Academy Sciences 2011;108:18554–9.
Agency, https://www.epa.gov/ghgemissions/understanding-global-warming [71] Carbis Bay G7 Summit Communique (June 15, 2021), https://www.whitehouse.
-potentials (accessed June 5, 2021). gov/briefing-room/statements-releases/2021/06/13/carbis-bay-g7-summit-co
[48] Foote E. Circumstances affecting the heat of the suns’s rays. The American mmunique/.
Journal of Science and Arts 1856;XXII:382–3. 2nd seriesNov. 1856. [72] Chu S. Carbon capture and sequestration. Science 2009;325:1599.
[49] Crowley TJ. Causes of climate change over the past 1000 years. Science 2000; [73] Majumdar A, Deutch J. Research opportunities for CO2 utilization and negative
289:270–7. emissions at the gigatonne scale. Joule 2018;2:805–9.
[50] Mann ME, Bradley RS, Hughes MK. Global-scale temperature patterns and climate [74] Emissions gap report 2019. UN Environment Programme; 26 November 2019.
forcing over the past six centuries. Nature 1998;392:779–87. https://www.unenvironment.org/resources/emissions-gap-report-2019.
[51] Canadell JG, Le Quere C, Raupach MR, Field CB, Buitenhuis ET, Ciais P, [75] U.S. Department of Commerce, National Oceanic and Atmospheric
Conway TJ, Gillett NP, Houghton RA, Marland G. Contributions to accelerating Administration, Carbon Tracker CT 2019, https://www.esrl.noaa.gov/gmd/ccgg/
atmospheric CO2 growth from economic activity, carbon intensity, and efficiency carbontracker/ (accessed April 28, 2020).
of natural sinks. Proceed. National Acad. Sci. 2007;104:18866–70. [76] Wise M, Calvin K, Thomson A, Clarke L, Bond-Lamberty B, Sands R, Smith SJ,
Janetos A, Edmonds J. Implications of limiting CO2 concentrations for land use
and energy. Science 2009;324:1183–6.

50
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

[77] Wilcox J. Carbon capture. New York: Springer-Verlag Publishing; March 2012. [105] Wilcox J, Haghapanah R, Rupp EC, He J, Lee K. Advancing adsorption and
ISBN 978-1-4614-2214-3. membrane separation processes for the gigaton carbon capture challenge. Annu
[78] Yang H, Xu Z, Fan M, Gupta R, Slimane RB, Bland AE, Wright I. Progress in Rev Chem Biomol Eng 2014;5:479–505.
carbon dioxide separation and capture: a review. J. Environ. Sci. 2008;20:14–27. [106] House KZ, Harvey CF, Aziz MJ, Schrag DP. The energy penalty of post-combustion
[79] Markewitz P, Kuckshinrichs W, Leitner W, Linssen J, Zapp P, Bongartz R, CO2 capture and storage and its implications for retrofitting the U.S. installed
Schreiber A, Muller TE. Worldwide innovations in the development of carbon base. Energy Environ. Sci. 2009;2:193–205.
capture technologies and the utilization of CO2. Energy Environ. Sci. 2012;5: [107] Gür TM, Homel M, Virkar AV. High performance solid oxide fuel cell operating on
7281–305. dry gasified coal. J. Power Sources 2010;195:1085–90.
[80] Boot-Handford ME, Abanades JC, Anthony EJ, Blunt MJ, Brandani S, Mac [108] Lee AC, Mitchell RE, Gür TM. Thermodynamic Analysis of Gasification-driven
Dowell N, Fernandez JR, Ferrari M-C, Gross R, Halleet JP, Haszeldine RS, Direct Carbon Fuel Cells. J. Power Sources 2009;194:774–85.
Heptonstall P, Lyngfelt A, Makuch Z, Mangano E, Porter RTJ, Pourkashanian M, [109] Gür TM. Comprehensive Review of Methane Conversion in Solid Oxide Fuel Cells:
Rochelle GT, Shah N, Yao JY, Fennell PS. Carbon capture and storage update. prospects for Efficient Electricity Generation from Natural Gas. Prog Energy
Energy Environ. Sci 2014;7:130–89. Combust Sci 2016;54:1–64.
[81] IPCC-AR5, Climate change 2014: synthesis report, https://www.ipcc.ch/site/ass [110] Vora SD, Lundberg WL, Pierre JF. Overview of U.S. DOE, Office of Fossil Energy’s
ets/uploads/2018/02/SYR_AR5_FINAL_full.pdf. Solid Oxide Fuel Cell Program. ECS Trans 2017;78(1):3–19.
[82] Trends in atmospheric CO2, mauna loa, hawaii. Global Monitoring Laboratory, [111] Wu F, Argyle MD, Dellenback PA, Fan M. Progress in O2 separation for oxy-fuel
Earth Systems Research Laboratories, U.S. Department of Commerce, National combustion – A promising way for cost-effective CO2 capture: a review. Progress
Oceanic & Atmospheric Administration; 2021. https://gml.noaa.gov/ccgg/trend Energy Combust. Sci 2018;67:188–205.
s/gr.html. accessed May 26. [112] Fu C, Gundersen T. Using exergy analysis to reduce power consumption in air
[83] Climate.gov, U.S. Department of Commerce, National Oceanic and Atmospheric separation units for oxy-combustion processes. Energy 2012;44:60–8.
Administration, https://www.climate.gov/news-features/understanding-climate [113] Castillo R. Thermodynamic analysis of a hard coal oxyfuel power plant with high
/climate-change-atmospheric-carbon-dioxide. temperature three-end membrane for air separation. Appl. Energy 2011;88:
[84] Connaughton J.L., Energy and Climate Policy, December 2007, http://belfercen 1480–93.
ter.ksg.harvard.edu/files/2007-12-12%20Connaughton%20Presentation%20 [114] Grand View Research, “Industrial Gases Market Size, Share & Trends Analysis
FINAL.pdf (accessed May 7, 2020). Report By Product (Nitrogen, Oxygen), By Application (Healthcare,
[85] International Energy Outlook 2019: with projections to 2050, Sept. 2019, Manufacturing), By Distribution (Onsite, Bulk), By Region, And Segment
#IEO2019, U.S. Energy Information Administration, https://www.eia.gov/outl Forecasts, 2020 –2027”, May 2020, https://www.grandviewresearch.com/indus
ooks/ieo/. try-analysis/industrial-gases-market.
[86] World energy outlook. International Energy Agency; 2016. http://www.iea.org/ [115] Smith AR, Klosek J. A review of air separation technologies and their integration
newsroom/news/2016/november/world-energy-outlook-2016.html. with energy conversion processes. Fuel 2001;70:115–34.
[87] U.S. Department of Energy, National Energy Technology Laboratory, Current and [116] Darde A, Prabhakar R, Tranier J-P, Perrin N. Air separation and flue gas
future technologies for gasification-based power generation, Volume 2: a compression and purification units for oxy-coal combustion systems. Energy
pathway study focused on carbon capture advanced power systems R&D using Procedia 2009;1:527–34.
bituminous coal, November 2010, DOE/NETL-2009/1389. [117] Pulverized Coal Oxycombustion Power Plants. Volume 1: bituminous coal to
[88] Vora SD. Office of Fossil Energy’s Solid Oxide Fuel Cell Program Overview. In: electricity, final report, publication no. doe/netl 2007/1291. Pittsburgh, PA: U.S.
15th Annual SECA Workshop; 2014. July 22–23, www.netl.doe.gov/File% Department of Energy, Office of Fossil Energy, National Energy Technology
20Library/Events/2014/2014%20SECA%20workshop/Shailesh-Vora.pdf. Laboratory; Aug. 2008.
accessed Nov. 27, 2020. [118] Aneke M, Wang M. Process analysis of pressurized oxy-coal power cycle for
[89] Pera-Titus M. Porous inorganic membranes for CO2 capture: present and carbon capture application integrated with liquid air power generation and
prospects. Chem. Rev. 2014;114:1413–92. binary cycle engines. Appl. Energy 2015;154:556–66.
[90] IPCC. Special report on carbon dioxide capture and storage. New York: [119] Yang RT. Gas separation by adsorption processes. Boston: Butterworth Publishers;
Cambridge University Press; 2005. ISBN: 92-9169-1135also, https://www.ipcc.ch 1987.
/site/assets/uploads/2018/03/srccs_chapter2-1.pdf. [120] Reid CR, O’Koye IP, Thomas KM. Adsorption of gases on carbon molecular sieves
[91] Capturing CO2, iea greenhouse gas r&d programme. International Energy Agency; used for air separation. Spherical adsorptives as probes for kinetic selectivity.
2007. ISBN: 978-1-898373-41-4. Langmuir 1998;14:2415–25.
[92] Rubin ES, Chen C, Rao AB. Cost and performance of fossil fuel power plants with [121] Sincar S, Golden TC, Rao MB. Activated carbon for gas separation and storage.
CO2 capture and storage. Energy Policy 2007;35:4444–54. Carbon N Y 1996;34:1–12.
[93] Psarra PC, Comello S, Bains P, Charoensawadpong P, Reishelstein S, Wilcox J. [122] Sircar S. Pressure swing adsorption. Ind. Eng. Chem. Res. 2002;41:1389–92.
Carbon capture and utilization in the industrial sector. Environ. Sci. Technol. [123] Koros WJ, Mahajan R. Pushing the limits on possibilities for large scale gas
2017;51:11440–9. separation: which strategies? J. Membrane Sci. 2000;175:181–96.
[94] Cuellar-Franca RM, Azapagic A. Carbon capture, storage and utilization [124] Bernardo P, Drioli E, Golemme G. Membrane gas separation: a review/State of the
technologies: a critical analysis and comparison of their life cycle environmental art. Ind. Eng. Chem. Res. 2009;48:4638–63.
impacts. J. CO2 Utilization 2015;9:82–102. [125] Gür TM. Permselectivity in Zeolite Filled Polysulfone Gas Separation Membranes.
[95] Carbon capture and storage database. National Energy Technology Laboratory, J. Membrane Sci. 1994;93:283–9.
US Department of Energy; 2020. https://netl.doe.gov/coal/carbon-storage/world [126] Liang CZ, Chung T-S, Lai J-y. A review of polymeric composite membranes for gas
wide-ccs-database. accessed Dec 13. separation and energy production. Progress Polymer Sci 2019;97:101141.
[96] Large Scale Carbon Capture Projects Database. CO2RE facilities database. Global [127] Gür TM, Belzner A, Huggins RA. A new class of oxygen selective chemically
CCS Institute; 2018. p. 1972–2029. https://co2re.co/FacilityData. accessed Dec. driven nonporous ceramic membranes: part 1. A-site doped perovskites.
13, 2020last update: Feb 9. J. Membrane Sci. 1992;75:151–62.
[97] Large Scale Carbon Capture Projects Database (last update: Feb 9, 2018), [128] Belzner A, Gür TM, Huggins RA. Oxygen Chemical Diffusion in Strontium Doped
1972–2029, https://datasource.kapsarc.org/explore/dataset/large-scale-car Lanthanum Manganites. Solid State Ionics 1992;57:327–37.
bon-capture-projects-database/information/?disjunctive.country&disjunctive.pr [129] Sunarso J, Hashim SS, Zhu N, Zhou W. Perovskite oxides applications in high
oject_name&disjunctive.project_lifecycle_stage&disjunctive.industry&disjunct temperature oxygen separation, solid oxide fuel cell and membrane reactor: a
ive.capture_type&disjunctive.primary_storage_type. review. Progress Energy Combust. Sci. 2017;61:57–77.
[98] KNOEMA, https://knoema.com/WCCSLSCC2016/large-scale-carbon-capture-pro [130] Chen L, SZ Yong, Ghoniem AF. Review: oxy-fuel combustion of pulverized coal:
jects-database-1972-2029 (accessed Dec. 13, 2020). characterization, fundamentals, stabilization and CFD modeling. Progress Energy
[99] CO2RE facilities. Global CCS Institute; 2020. https://co2re.co/FacilityData. Combust. Sci. 2012;38:156–214.
accessed Dec.23. [131] Scheffknecht G, Al-Makhadmed L, Schnell U, Maier J. Oxy-fuel coal combustion –
[100] Global CCS Institute, New wave of CCS activity: ten large-scale projects A review of the state-of-the-art. Intern. J. Greenhouse Gas Control 2011;55:
announced, Oct. 29, 2019, https://www.globalccsinstitute.com/news-media/pr 516–35.
ess-room/media-releases/new-wave-of-ccs-activity-ten-large-scale-projects-anno [132] Zhang Z, Li X, Luo C, Zhang L, Xu Y, Wu Y, Liu J, Duam Y, Zheng C. Investigation
unced/. on the thermodynamic calculation of a 35 MWth oxy-fuel combustion coal-fired
[101] Gür TM. Review of electrical energy storage technologies, materials, and systems: boiler. Int. J. Greenhouse Gas control 2018;71:36–45.
challenges and prospects for large-scale grid storage. Energy Environ. Sci. 2018; [133] Toftegaard MB, Brix J, Jensen PA, Glarborg P, Jensen AD. Oxy-fuel combustion of
11:2696–767. solid fuels. Progress Energy Combust. Sci. 2010;36:581–625.
[102] U.S. Department Of Energy, Office of Fossil Energy, “Carbon Capture Program”, [134] Wall T, Liu Y, Spero C, Elliott L, Khare S, Rathnam R, Zeenathal F, Maghtaderi B,
https://www.energy.gov/fe/program-approach2nd-generation-technologies Buhre B, Sheng C, Gupta R, Yamada T, Makino K, Yu J. An overview of oxyfuel
(accessed May 7, 2020). coal combustion – State of the art research and technology development. Chem.
[103] House KZ, Baclig AC, Rajan M, van Nierop EA, Wilcox J, Herzog HJ. Economic Eng. Res. Design 2009;87:1003–16.
and energetic analysis of capturing CO2 from ambient air. Proceedings of the [135] U.S. DOE- National Energy Technology Laboratory, Oxy-combustion. https://netl.
National Academy of Sciences (PNAS) 2011;108:20428–33. doe.gov/node/7477 (accessed June 20, 2021).
[104] Lackner KS, Brennan S, Matter JM, Park A-HA, Wright A, van der Zwaan B. The [136] Stanger R, Wall T. Sulphur impacts during pulverized coal combustion in oxy-fuel
urgency of the development of CO2 capture from ambient air. Proceed. National technology for carbon capture and storage. Progress Energy Combust. Sci. 2011;
Acad. Sci. (PNAS) 2012;109:13156–62. 37:69–88.
[137] Tan Y, Croiset E, Douglas MA, Thambimuthu KV. Combustion characteristics of
coal in a mixture of oxygen and recycled flue gas. Fuel 2006;85:507–12.

51
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

[138] Habib MA, Badr MH, Ahmed SF, Ben-Mansour R, Mezghani K, Imashuku S, la [171] Ippolito D, Deleebeeck L, Hansen KK. Effect of CeO2 addition on hybrid direct
O GJ, Shao-Horn Y, Mancini ND, Mitsos A, Kirchen P, Ghoneim AF. A review of carbon fuel cell performance. J. Electrochem. Soc. 2017;164:F328.
recent developments in carbon capture utilizing oxy-fuel combustion in [172] Wu H, Xiao J, Zeng X, Li X, Yang J, Zou Y, Liu S, Dong P, Zhang Y, Liu J. A high
conventional and ion-transport membrane systems. Int. J. Energy Res. 2011;35: performance direct carbon solid oxide fuel cell – A green pathway for brown coal
741–64. utilization. Appl. Energy 2019;248:679–87.
[139] Fujimori T, Yamada T. Realization of oxyfuel combustion for near zero emission [173] Chen Q, Qiu Q, Yan X, Zhou M, Zhang Y, Liu Z, Cai W, Wang W, Liu J. A compact
power generation. Proceed. Combust. Inst. 2013;34:2111–30. and seal-less direct carbon solid oxide fuel cell stack stepping into practical
[140] Zeng Z, Natesan K, Cai Z, Gosztola DJ. Effects of chlorine in ash on the corrosion application. Appl. Energy 2020;278:115657.
performance of Ni-based alloys in a simulated oxy-fuel environment. Energy Fuels [174] Fuente-Cuesta A, Jiang C, Arenillas A, Irvine JTS. Role of coal characteristics in
2018;32:10502–12. the electrochemical behavior of hybrid direct carbon fuel cells. Energy Environ.
[141] Wall T, Stanger R, Liu Y. Gas cleaning challenges for coal-fired oxy-fuel Sci. 2016;9:2868–80.
technology with carbon capture and storage. Fuel 2013;108:85–90. [175] Gür TM. Low-carbon electricity is great: what about “less-carbon”?
[142] Ja’baz I, Chen J, Etschmann B, Ninomiya Y, Zhang L. High-temperature tube J. Electrochem. Soc. 2017;164:F1587–90.
corrosion upon the interaction with Victorian brown coal fly ash under the oxy- [176] Cherepy NJ, Krueger R, Fiet KJ, Jankowski AJ, Cooper JF. Direct conversion of
fuel combustion conditions. Proceed. Combust. Inst. 2017;36:3941–8. carbon in molten carbonate fuel cell. J. Electrochem. Soc. 2005;152:A80–7.
[143] Stanger R, Wall T, Sporl R, Paneru M, Grathwohl S, Weidmann M, Scheffknecht G, [177] Jiang C, Ma J, Bonaccorso AD, Irvine JTS. Demonstration of high power, direct
McDonald D, Myohanene K, Ritvanen J, Rahiala S, Hyppanen T, Mletzko J, conversion of waste-derived carbon in a hybrid direct carbon fuel cell. Energy
Kather J. Oxyfuel combustion for CO2 capture in power plants. Intern. J. Environ. Sci. 2012;5:6973–80.
Greenhouse Gas Control 2015;40:55–125. [178] Lee AC, Mitchell RE, Gür TM. Modeling of CO2 gasification of carbon for
[144] Yin C, Yan J. Oxy-fuel combustion of pulverized fuels: combustion fundamentals integration with solid oxide fuel cells. AIChE Journal 2009;55:983–92.
and modeling. Appl. Energy 2016;162:742–62. [179] Giddey S, Badwal SPS, Kulkarni A, Munnings C. A comprehensive review of direct
[145] Wall T, Stanger R, Santos S. Demonstrations of coal-fired oxy-fuel technology for carbon fuel cell technology. Progress Energy Combust. Sci. 2012;38:360–99.
carbon capture and storage and issues with commercial deployment. Intern. J. [180] Jiang C, Ma J, Corre G, Jain SL, Irvine JTS. Challenges in developing direct
Greenhouse Gas Control 2011;55:S5–15. carbon fuel cells. Chem. Soc. Rev. 2017;46:2889–912.
[146] Lyngfelt A. Chemical-looping combustion of solid fuels – Status and development. [181] Cao T, Huang K, Shi Y, Cai N. Recent advances in high-temperature carbon-air
Appl. Energy 2014;113:1869–73. fuel cells. Energy Environ. Sci. 2017;10:460–90.
[147] Hossain MM, de Lasa HI. Chemical-looping combustion (CLC) for inherent CO2 [182] Gür TM. Mechanistic Modes for Solid Carbon Conversion in High Temperature
separations – a review. Chem. Eng. Sci. 2008;63:4433–51. Fuel Cells. J. Electrochem. Soc. 2010;157:B571–759.
[148] Adanez J, Abad A, Gaecia-Labiano F, Gayan P, de Diego LF. Progress in chemical- [183] Sumida K, Rogow DL, Mason JA, McDonald TM, Bloch ED, Herm ZR, Bae T-H,
looping combustion and reforming technologies. Progress Energy Combust. Sci. Long JR. Carbon dioxide capture in metal-organic frameworks. Chem. Rev. 2012;
2012;38:215–82. 112:724–81.
[149] Goldstein EA, Gür TM, Mitchell RE. Modeling defect transport during Cu [184] Yu C-H, Huang C-H, Tan C-S. A review of CO2 capture by absorption and
oxidation. Corrosion Sci. 2015;99:53–65. adsorption. Aerosol Air Quality Res 2012;12:745–69.
[150] U.S. DOE- National Energy Technology Laboratory, Chemical Looping [185] Pennline HW, Luebke DR, Jones KL, Myers CR, Morsi BI, Heintz YJ, Ilconich JB.
Combustion https://www.netl.doe.gov/node/7478 (accessed June 20, 2021). Progress in carbon dioxide capture and separation research for gasification-based
[151] Luo M, Yi Y, Wang C, Liu K, Pan J, Wang Q. Energy and exergy analysis of power power generation point sources. Fuel Process. Technol. 2008;89:897–907.
generation systems with chemical looping combustion of coal. Chem. Eng. [186] Figueroa JD, Fout T, Plasynski S, McIlvried H. Advances in CO2 capture
Technol. 2018;41:776–87. technology – The U.S. Department of Energy’s carbon sequestration program.
[152] Cho P, Mattisson T, Lyngfelt A. Comparison of iron-, nickel-, copper- and Intern. J. Greenhouse Gas Control 2008;2:9–20.
manganese-based oxygen carriers for chemical looping combustion. Fuel 2004; [187] de Coninck H, Benson SM. Carbon dioxide capture and storage: issues and
83:1215–25. prospects. Ann. Rev. Environ. Resour. 2014;39:243–70.
[153] Adanez J, de Diego LF, Gaecia-Labiano F, Gayan P, Abad A, Palacios JM. Selection [188] Lackner KS. A guide to CO2 sequestration. Science 2003;300:1677–8.
of oxygen carriers for chemical-looping combustion. Energy Fuels 2004;18:371–7. [189] Leung DYC, Caramanna G, Maroto-Valer MM. An overview of current status of
[154] Matzen M, Pinkerton J, Wang X, Demirel Y. Use of natural ores as oxygen carriers carbon dioxide capture and storage technologies. Renew. Sustain. Energy Reviews
in chemical looping combustion: a review. Intern. J. Greenhouse Gas Control 2014;39:426–43.
2017;65:1–14. [190] Samanta A, Zhao A, Shimizu GKH, Sarkar P, Gupta R. Post-combustion CO2
[155] Nisancioglu K, Gür TM. Potentiostatic step technique to study ionic transport in capture using solid sorbents: a review. Ind. Eng. Chem. Res. 2012;51:1438–63.
mixed conductors. Solid State Ionics 1994;72:199–203. [191] Abu-Zahra MRM, Niederer JPM, Feron PHM, Versteeg GF. CO2 capture from
[156] Readman JE, Olafsen A, Larring Y, Blom R. La0.8Sr0.2Co0.2Fe0.8 O3-d as a potential power plants Part II: a parametric study of the economical performance based on
oxygen carrier in a chemical looping type reactor, an in-situ powder X-ray mono-ethanolamine. Intern. J. Greenhouse Gas Control 2007;1:135–42.
diffraction study. J. Mater. Chem. 2005;15:1931–7. [192] Yamasaki A. An overview of CO2 mitigation options for global warming –
[157] Lyngfelt A. Chemical Looping Combustion: status and development challenges. Emphasizing CO2 sequestration options. J. Chem. Engr. Japan 2003;36:361–75.
Energy Fuels 2020;34:9077–93. [193] Luis P. Use of monoethanolamine (MEA) for CO2 capture in a global scenario:
[158] MIT Technology Review July 5, 2016, https://www.technologyreview.com/ consequences and alternatives. Desalination 2016;380:93–9.
2016/07/05/158978/a-mississippi-power-plant-highlights-all-thats-wrong-with- [194] Strait R, Nagveka M. Carbon dioxide capture and storage in the nitrogen and
clean-coal/(accessed Dec. 12, 2020). syngas industries. Nitrogen Syngas 2010;303:16–9.
[159] Clean Coal Power Initiative, https://netl.doe.gov/node/2302 (accessed Dec. 12, [195] Rochelle GT. Amine scrubbing for CO2 capture. Science 2009;325:1652–4.
2020). [196] Idem R, Supap T, Shi H, Gelowitz D, Ball M, Campbell C, Tontiwachwuthikul P.
[160] International Energy Agency, Paris, July 7, 2017, https://www.iea.org/comment Practical experience in post-combustion CO2 capture using reactive solvents in
aries/we-cant-let-kemper-slow-the-progress-of-carbon-capture-and-storage large pilot and demonstration plants. Intern. J. Greenhouse Gas Control 2015;40:
(accessed Dec 12, 2020). 6–25.
[161] Higman C, Tam S. Advances in coal gasification, hydrogenation, and gas treating [197] Walters MS, Edgar TF, Rochelle GT. Regulatory control of amine scrubbing for
for the production of chemicals and fuels. Chem. Rev. 2014;114:1673–708. CO2 capture from power plants. Ind. Eng. Chem. Res. 2016;55:4646–57.
[162] U.S. Department of Energy, http://energy.gov/fe/how-coal-gasification-power-pl [198] El Hadri N, Quang DV, Goetheer ELV, Abu Zahra MRM. Aqueous amine solution
ants-work. characterization for post-combustion CO2 capture process. Appl. Energy 2017;
[163] Zahradnik RL, Elikan L, Archer DH. A coal-burning solid-electrolyte fuel cell 185:1433–49.
power plant. Am. Chem. Soc. Div. Fuel Chem. 1964;8:212–28. [199] Rao AB, Rubin ES. A technical, economic, and environmental assessment of
[164] Archer DH, Zahradnik RL. Design of a 100-kilowatt coal-burning fuel cell power amine-based CO2 capture technology for power plant greenhouse gas control.
system. Am. Chem. Soc. Div. Fuel Chem. 1967;11:107–46. Environ. Sci. Technol. 2002;36:4467–75.
[165] Zahradnik RL, Elikan L, Archer DH. A coal-burning solid-electrolyte fuel cell [200] Nielsen CJ, Herrmann H, Weller C. Atmospheric chemistry and environmental
power plant. Fuel Cell Systems, Chapter 25, Advances in Chemistry 1969;47: impact of the use of amines in carbon capture and storage (CCS). Chem. Soc. Rev.
343–56. 2012;41:6684–704.
[166] Weaver R.D., Tietz L., Cubicciotti D. Direct Use of Coal in a Fuel Cell: feasibility [201] da Silva EF, Booth AM. Emissions from postcombustion CO2 capture plants.
Investigation, Final Report, EPA-68-02-1808 (June 1, 1975). Environ. Sci. Technol. 2013;47:659–60.
[167] Weaver RD, Leach SC, Bayce AE, Nanis L. Direct electrochemical generation of [202] Reynolds AJ, Verheyen TV, Adeloju SB, Meuleman E, Feron P. Towards
electricity from coal. Menlo Park, CA: SRI; 1979. p. 94025. SAN-0115/105-1. commercial scale postcombustion capture of CO2 with monoethanolamine
[168] Weaver RD, Yasuda M, Bayce AE, Nanis L. Direct Electrochemical Generation of solvent: key considerations for solvent management and environmental impacts.
Electricity from Coal. In: Annual Report to U.S. Energy Research and Environ. Sci. Technol. 2012;46:3643–54.
Development Administration for the period January 1976 to February 1977. [203] Li J, Chen L, Ye Y, Qi Z. Solubility of CO2 in the mixed solvents system of
Menlo Park, CA: Stanford Research Institute; May 1977. alkolamines and poly(ethylene) glycol 200. J. Chem. Eng. Data 2014;59:1781–7.
[169] U.S. Department of Energy, National Energy Technology Laboratory (NETL), [204] Li HI, Yan JY, Campana PE. Feasibility of integrating solar energy into a power
https://netl.doe.gov/coal/fuel-cells. plant with amine-based chemical absorption for CO2 capture. Intern. J.
[170] Vora SD, Jesionowski G, Williams MC. Overview of the U.S. Department of Energy Greenhouse gas Control 2012;9:272–80.
Office of Fossil Energy’s Solid Oxide Fuel Cell Program for FY2019. ECS Trans [205] Jande YAC, Asif M, Shim SM, Kim WS. Energy minimization in
2019;91(1):27–39. monoethanolamine-based CO2 capture using capacitive deionization. Intern. J.
Energy Res. 2014;38:1531–40.

52
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

[206] Yeh JT, Resnik KP, Rygle K, Pennline HW. Semi-batch absorption and [240] Hoffman JS, Pennline HW. Study of regenerable sorbents for CO2 capture. In:
regeneration studies for CO2 capture by aqueous ammonia. Fuel Process. Technol. Proceedings of First National Conference on carbon sequestration; May 2001.
2005;86:1533–46. [241] Reijnen K, van Brakel J. Gas Cleaning at High Temperatures and High Pressures: a
[207] Bandyopadhyay A. Amine versus ammonia absorption of CO2 as a measure of Review. Powder Technol 1984;40:81–111.
reducing GHG emissions: a critical analysis. Clean Techn. Environ. Policy 2011; [242] Kremer J, Galloy A, Strohle J, Epple B. Continuous CO2 capture in a 1-MW
13:269–94. carbonate looping pilot plant. Chem. Eng. Technol. 2013;36:1518–24.
[208] Huang HP, Shi Y, Li W, Chang SG. Dual alkali approaches for the capture and [243] Liu W, An H, Qin C, Yin J, Wang G, Feng B, Xu M. Performance enhancement of
separation of CO2. Energy Fuels 2001;15:263–8. calcium oxide sorbents for cyclic CO2 capture – A review. Energy Fuels 2012;26:
[209] Erisman JW, Sutton MA, Galloway J, Klimont Z, Winiwarter W. How a century of 2751–67.
ammonia synthesis changed the world. Nature Geosci 2008;1:636–9. [244] Fauth DJ, Frommell EA, Hoffman JS, Reasbeck RP, Pennline HW. Eutectic salt
[210] Knuutila H, Svendsen HF, Anttila M. CO2 capture from coal-fired power plants promoted lithium zirconate: novel high temperature sorbent for CO2 capture. Fuel
based on sodium carbonate slurry: a systems feasibility and sensitivity study. Process. Technol. 2005;86:1503–21.
Intern. J. Greenhouse Gas Control 2009;3:143–51. [245] Kato M, Nakagawa K, Esaki K, Maezawa Y, Takeda S, Kogo R, Hagiwara Y. Novel
[211] Angell CA, Ansari Y, Zhao Z. Ionic liquids: past, present and future. Faraday CO2 absorbents using lithium containing oxides. Int. J. Appl. Ceram. Technol.
Discuss 2012;154:9–27. 2005;2:467–75.
[212] Yu G, Zhao D, Wen L, Yang S, Chen X. Viscosity of ionic liquids: database, [246] Ortiz AL, Bretado MAE, Velderrain VG, Zarazoga MM. Experimental and
observation, and quantitative structure-property relationship analysis. AIChE modeling kinetic study of the CO2 absorption by Li4SiO4. Int. J. Hydrogen Energy
Journal 2012;58:2885–99. 2014;39:16656–66.
[213] Zhang S, Sun N, He X, Lu X, Zhang X. Physical properties of ionic liquids: database [247] Liang J, Liang Z, Zou R, Zhao Y. Heterogeneous catalysis in zeolites, mesoporous
and evaluation. J. Phys. Chem. Ref. Data 2006;35:1475–517. silica, and metal-organic frameworks. Adv. Mater. 2017;29:1701139.
[214] Armand M, Endres F, MacFarlane DR, Ohno H, Scrosati B. Ionic-liquid materials [248] U.S. National Science Foundation, https://nsf.gov/news/mmg/mmg_disp.jsp?
for the electrochemical challenges of the future. Nature Mater 2009;8:621–9. med_id=70939&from= (accessed June 10, 2021).
[215] Olivier-Bourbigou H, Magna L. Ionic liquids: perspectives for organic and [249] Harlick PJE, Tezel FH. An experimental adsorbent screening study of CO2 removal
catalytic reactions. J. Molecular Catal. A: Chemical 2002;182-183:419–37. from N2. Microposrous Mesoporous Mater 2004;76:71–9.
[216] Ramdin M, de Loos TW, Vlugt TJH. State-of-the-art of CO2 capture with ionic [250] Database of Zeolite Structures, Structure Commission of the International Zeolite
liquids. Ind. Eng. Chem. Res. 2012;51:8149–77. Association (IZA-SC), http://www.iza-structure.org/databases/(accessed Jan. 5,
[217] Cadena C, Anthony JL, Shah JK, Morrow TI, Brennecke JF, Maginn EJ. Why is 2021).
CO2 soluble in imidazolium-based ionic liquids? J. Am. Chem. Soc. 2004;126: [251] Tao Y, Kanoh H, Abrams L, Kaneko K. Mesopore-modified zeolites: preparation,
5300–8. characterization and applications. Chem. Rev. 2006;106:896–910.
[218] Kasahara S, Kamio E, Ishigami T, Matsuyama H. Amino acid ionic liquid-based [252] Kumar S, Srivastava R, Koh J. Utilization of zeolites as CO2 capturing agents:
facilitated transport membranes for CO2 separation. Chem. Commun. 2012;48: advances and future perspectives. J. CO2 Utilization 2020;41:101251.
6903–5. [253] Cheung O, Hedin N. Zeolites and related sorbents with narrow pores for CO2
[219] Palomar J, Gonzalez-Miquel M, Polo A, Rodriguez F. Understanding the physical separation from flue gas. RSC Adv 2014;4:14480–94.
absorption of CO2 in ionic liquids using the COSMOS-RS method. Ind. Eng. Chem. [254] West RC. CRC handbook of chemistry and physics. CRC Press; 1967.
Res. 2011;50:3452–63. [255] Walton KS, Abney MB, LeVan MD. CO2 adsorption in Y and X zeolites modified by
[220] Carvalho PJ, Coutinho JAP. On the nonideality of CO2 solutions in ionic liquids alkali metal cation exchange. Microporous Mesoporous Mater 2006;91:78–84.
and other low volatile solvents. J. Phys. Chem. Lett. 2010;1:774–80. [256] Chue KT, Kim JN, Yoo YJ, Cho SH, Yang RT. Comparison of activated carbon and
[221] Bates ED, Mayton RD, Ntai I, Davis JH. CO2 capture by a task-specific ionic liquid. zeolite 13X for CO2 recovery from flue gas by pressure swing adsorption. Ind. Eng.
J. Am. Chem. Soc. 2002;124:926–7. Chem. Res. 1995;34:591–8.
[222] Hu H, Li F, Xia Q, Liao I, Fan M. Research on influencing factors and mechanism [257] O’Keefe M, Yaghi OM. Deconstructing the crystal structure of metal-organic
of CO2 absorption by poly-amino-based ionic liquids. Intern. J. Greenhouse Gas frameworks and related materials into their underlying nets. Chem. Rev. 2012;
Control 2014;31:33–40. 112:675–702.
[223] Hu P, Zhang R, Liu Z, Liu H, Xu C, Mang X, Liang M, Liang S. Absorption [258] Furukawa H, Cordova KE, O’Keeffe M, Yaghi OM. The chemistry and applications
performance and mechanism of CO2 in aqueous solutions of amine-based ionic of metal-organic frameworks. Science 2013;341:6149.
liquids. Energy Fuels 2015;29:6019–24. [259] Wang Q, Astruc D. State of the art and prospects in metal-organic framework
[224] Choi S, Drese JH, Jones CW. Adsorbent materials for carbon dioxide capture from (MOF)-based and MOF-derived nanocatalysis. Chem. Rev. 2020;120:1438–511.
large anthropogenic point sources. ChemSusChem 2009;2:796–854. [260] Rogge SMJ, Bavykina A, Hajek J, Garcia H, Olivos-Suarez AI, Sepulveda-
[225] Chaffee AL, Knowles GP, Liang Z, Zhang J, Xiao P, Webley PA. CO2 capture by Escribano A, Vimont A, Clet G, Bazin P, Kapteijn F, Daturi M, Ramos-
adsorption: materials and process development. Int. J. Greenhouse Gas Control Fernandez EV, Liabres I Xamena FX, Van Speybroeck V, Gascon J. Metal-organic
2007;1:11–8. and covalent frameworks as single-site catalysis. Chem. Soc. Rev. 2017;46:
[226] Computational Chemistry Comparison and Benchmark DataBase, National 3134–84.
Institute of Standards and Technology (Release 21 (August 2020), Standard [261] Lin Y, Kong C, Zhang Q, Chen L. Metal-organic frameworks for carbon dioxide
Reference Database 101), https://cccbdb.nist.gov/quadlistx.asp (accessed Jan. 4, capture and methane storage. Adv. Energy Mater. 2017;7:1601296.
2021). [262] Zhao Z, Ma X, Kasik A, Li Z, Lin YS. Gas separation properties of metal organic
[227] Krishnamurthy S, Rao VR, Guntuka S, Sharratt P, Haghpanah R, Rajendran A, framework (MOF-5) membranes. Ind. Eng. Chem. Res. 2013;52:1102–8.
Amanullah M, Karimi IA, Farooq S. CO2 capture from dry flue gas by vacuum [263] Li J-R, Kuppler RJ, Zhou H-C. Selective gas adsorption and separation in metal-
swing adsorption: a pilot plant study. AIChE Journal 2014;60:1830–42. organic frameworks. Chem. Soc. Rev. 2009;38:1477–504.
[228] Clausse M, Merel J, Meunier F. Numerical parametric study on CO2 capture by [264] Danaci D, Bui M, Mac Dowell N, Petit C. Exploring the limits of adsorption-based
indirect thermal swing adsorption. Int. J. Greenhouse Gas Control 2011;5: CO2 capture using MOFs with PVSA – from molecular design to process
1206–13. economics. Mol. Syst. Des. Eng. 2020;5:212–31.
[229] Kulkarni AR, Sholl DS. Analysis of equilibrium-based TSA processes for direct [265] Schnobrich JK, Koh K, Sura KN, Matzger AJ. A framework for predicting surface
capture of CO2 from air. Ind. Eng. Chem. Res. 2012;51:8631–45. areas in microporous coordination polymers. Langmuir 2010;26:5808.
[230] Yue MB, Sun LB, Cao Y, Wang Y, Wang ZJ, Zhu JH. Efficient CO2 capturer derived [266] Liu J, Thallapally PK, McGrail BP, Brown DR, Lio J. Progress in adsorption-based
from as-synthesized MCM-41 modified with amine. Chem. Eur. J. 2008;14: CO2 capture by metal-organic frameworks. Chem. Soc. Rev. 2012;41. 23-8-22.
3442–51. [267] Banerjee R, Furukawa H, Britt D, Knobler C, O’Keeffe M, Yaghi OM. Control of
[231] Jiang L, Gonzalez-Diaz A, Ling-Chin J, Roskilly AP, Smallbone AJ. Post- pore size and functionality in isoreticular zeolitic ididazolate frameworks and
combustion CO2 capture from a natural gas combined cycle power plant using their carbon dioxide selective capture properties. J. Am. Chem. Soc. 2009;131:
activated carbon adsorption. Appl. Energy 2019;245:1–15. 3875–7.
[232] Sevilla M, Parra JB, Fuertes AB. Assessment of the role of micropore size and N- [268] Liu J, Tian J, Thallapally PK, McGrail BP. Selective CO2 capture from flue gas
doping in CO2 capture by porous carbons. Appl. Mater. Interfaces 2013;5:6360–8. using metal-organic frameworks – A fixed bed study. J. Phys. Chem. C 2012;116:
[233] Gür TM, Bent SF, Prinz FB. Nanostructuring Materials for Solar-to-Hydrogen 9575–81.
Conversion. J. Phys. Chem. C 2014;118:21301–15. [269] Furukawa H, Ko N, Go YB, Aratani N, Choi SB, Choi E, Yazaydin AO, Snurr RQ,
[234] Gür TM. Tools and Strategies for Developing New Materials. Adv. Mater. 1996;8: O’Keeffe M, Kim J, Yaghi OM. Ultrahigh porosity in metal-organic frameworks.
883. Science 2010;329:424–8.
[235] Shen J, Liu G, Huang K, Jin W, Lee K-R, Xu N. Membranes with fast and selective [270] Stadler H, Beggel F, Habermehl M, Persigehl B, Kneer R, Modigell M, Jeschke P.
gas-transport channels of laminar grapheme oxide for efficient CO2 capture. Oxyfuel coal combustion by efficient integration of oxygen transport membranes.
Angew. Chem. 2015;127:588–92. Intern. J. Greenhouse Gas Control 2017;5:7–15.
[236] Liu Y, Wilcox J. Molecular simulation studies of CO2 adsorption by carbon model [271] Lonsdale HK. The growth of membrane technology. J. Membrane Sci. 1982;10:
compounds for carbon capture and sequestration applications. Environ. Sci. 81–181.
Technol. 2013;1:95–101. [272] Yegani R, Hirozawa H, Teramoto M, Himei H, Okada O, Takigawa T, Ohmura N,
[237] Özdemir E, Schroeder K. Effect of moisture on adsorption isotherms and Matsumiya N, Matsumaya H. Selective separation of CO2 by using novel
adsorption capacities of CO2 on coals. Energy Fuels 2009;5:2821–31. facilitated transport membrane at elevated temperatures and pressures.
[238] Gholidoust A, Atkinson JD, Hashisho Z. Enhancing CO2 adsorption via amine- J. membrane Sci. 2007;291:157–64.
impregnated activated carbon from oil sands coke. Energy Fuels 2017;2:1754–63. [273] Favre E. Carbon dioxide recovery from post-combustion processes: can gas
[239] Chester AW, Derouane EG, editors. Zeolite characterization and catalysis: a permeation membranes compete with absorption? J. Membrane Sci. 2007;294:
tutorial. New York: Springer; 2009. 50–9.

53
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

[274] Verweij H. Inorganic membranes. Current Opinion Chem. Eng. 2012;1:156–62. [309] McCoy ST, Rubin ES. An engineering-economic model of pipeline transport of
[275] Brunetti A, Scura F, Barbieri G, Drioli E. Membrane technologies for CO2 CO2 with application to carbon capture and storage. Intern. J. Greenhouse Gas
separation. J. Membrane Sci. 2010;359:115–25. Control 2008;2:219–29.
[276] Anderson M, Wang H, Lin YS. Inorganic membranes for carbon dioxide and [310] Gao I, Fang M, Li H, Hetland J. Cost analysis of CO2 transportation: case study in
nitrogen separation. Reviews Chem. Eng. 2012;28:101–21. China. Energy Procedia 2011;4:5974–81.
[277] Yang H-C, Hou J, Chen V, Xu Z-K. Surface and interface engineering of organic- [311] Forbes SM, Verma P, Curry TE, Friedmann SJ, Wade SM. Guidelines for carbon
inorganic composite membranes. J. Mater. Chem. A 2016;4:9716–29. dioxide capture, transport, and storage. World Resource Institute; 2008. No.
[278] Ostwal M, Singh RP, Dec SF, Lusk MT, Way JD. 3-Aminopropyltriethoxysilane 20113082154 p.144.
functionalized inorganic membranes for high temperature CO2/N2 separation. [312] Metz B, Davidson O, de Coninck H, Loos M, Meyer L. IPCC special report on
J. Membrane Sci. 2011;369:139–47. carbon dioxide capture and storage, intergovernmental panel on climate change.
[279] Lei L, Bai L, Lindbrathen A, Pan F, Zhang X, He X. Carbon membranes for CO2 Geneva (Switzerland): Working group III; 2005. https://www.ipcc.ch/report
removal: status and perspectives from materials to processes. Chem. Eng. J. 2020; /carbon-dioxide-capture-and-storage/. accessed June 5, 2021.
401:126084. [313] Aminu MD, Nabavi SA, Roichelle CA, Manovic V. A review of developments in
[280] Lu C, Bai H, Wu B, Su F, Hwang JF. Comparative study of CO2 capture by carbon carbon dioxide storage. Appl. Energy 2017;208:1389–419.
nanotubes, activated carbon, and zeolites. Energy Fuels 2008;22:3050–6. [314] Benson SM, Cole DR. CO2 sequestration in deep sedimentary formations. Elements
[281] Alcaniz-Monge J, Marco-Lazar JP, Lillo-Rodenas MA. CO2 separation by carbon 2008;4:325–31.
molecular sieve monoliths prepared from nitrated coal tar pitch. Fuel Process. [315] Zoback MD, Gorelick SM. Earthquake triggering and large-scale geologic storage
Technol. 2011;92:915–9. of carbon dioxide. Proceed. National Acad. Sci. 2012;109:10164–8.
[282] Cinke M, Li J, Bauschlicher CW, Ricca A, Meyyappan M. CO2 adsorption on [316] Szulczewski ML, MacMinn CW, Herzog H, Juanes R. Lifetime of carbon capture
single-walled carbon nanotubes. Chem. Phys. Lett. 2003;376:761–6. and storage as a climate-change mitigation technology. Proceedings of National
[283] Sakamoto Y, Nagata K, Yogo K, Yamada K. Preparation and CO2 separation Academy of Sciences (PNAS) 2012;109:5185–9.
properties of amine-modified mesoporous silica membranes. Microporous [317] Kearns J, Teletzke G, Palmer J, Thomann H, Kheshgi H, Chen Y-HH, Paltsev S,
Mesoporous Mater 2007;101:303–11. Herzog H. Developing a consistent database for regional geologic CO2 storage
[284] Gottlicher G, Pruschek R. Comparison of CO2 removal systems for fossil fuelled capacity worldwide. Energy Procedia 2017;114:4697–709.
power plants. Energy Convers. Manag. 1997;38:S173–8. [318] Holdren JP, et al. Ending the energy stalemate: a bipartisan strategy to meet
[285] Dicks AL. Molten carbonate fuel cells. Current Opinion Solid State Mater. Sci. america’s energy challenges. The National Commission on Energy Policy; 2004.
2004;8:379–83. December 2004.
[286] Selman JR. Molten-salt fuel cells – Technical and economic challenges. J. Power [319] Benson SM, Bennaceur K, Cook P, Davison J, de Coninck HC, et al. Carbon dioxide
Sources 2006;160:852–7. capture and storage. In: Gomez-Echeverri L, Johnson TB, Nakicenovic N,
[287] Weaver JL, Winnick J. The molten carbonate carbon dioxide concentrator: Patwardhan A, editors. Global energy assessment: toward a sustainable future.
cathode performance at high CO2 utilization. J. Electrochem. Soc. 1983;130: Cambridge, UK/New York: Cambridge University Press; 2012. p. 993–1068.
20–8. [320] White CM, Strazisar BR, Granite EJ, Hoffman JS, Pennline HW. Separation and
[288] Chung SJ, Park JH, Li D, Ida J-I, Kumakiri I, Lin JYS. Dual-phase metal-carbonate capture of CO2 from large stationary sources and sequestration in geological
membrane for high-temperature carbon dioxide separation. Ind. Eng. Chem. Res. formations – coal beds and deep saline aquifers. J. Air Waste Manage. Assoc.
2005;44:7999–8006. (AWMA) 2003;53:645–715.
[289] Wade JL, Lackner KS, West AC. Transport model for a high temperature, mixed [321] Snaebjornsdottir SO, Sigfusson B, Marieni C, Goldberg D, Gislason SR. Carbon
conducting CO2 separation membrane. Solid State Ionics 2007;178:1530–40. dioxide storage through mineral carbonation. Nature Reviews Earth &
[290] Wade JL, Lee C, West AC, Lackner KS. Composite electrolyte membrane for high Environment 2020;1:90–102.
temperature CO2 separation. J. Membrane Sci. 2011;369:20–9. [322] Carbon Sequestration Technology Roadmap. Carbon sequestration leadership
[291] Li X, Huang K, Jin X. Mathematical modeling of high-temperature multiphase forum. U.S. Department of Energy, Office of Fossil Energy; 2017. https://www.
solid/molten carbonate membranes for CO2 capture. J. Electrochem. Soc. 2020; cslforum.org/cslf/sites/default/files/2017CSLFTechnologyRoadmap.pdf.
167. 164512 1-20. accessed June 6, 2021.
[292] Anderson M, Lin YS. Carbonate-ceramic dual-phase membrane for carbon dioxide [323] The Global Status of CCS Report 2020, Global CCS Institute, https://www.
separation. J. Membrane Sci. 2010;357:122–9. globalccsinstitute.com/resources/global-status-report/ (accessed Dec. 14, 2020).
[293] Yin H, Mao X, Tang D, Xiao W, Xing L, Zhu H, Wang D, Sadoway DR. Capture and [324] Global carbon capture and storage capacity grew by 33% in 2020, says new
electrochemical conversion of CO2 to value-added carbon and oxygen by molten report, S&P Platts, Dec. 1, 2020, https://www.spglobal.com/platts/en/ma
salt electrolysis. Energy Environ. Sci. 2013;6:1538–45. rket-insights/latest-news/coal/120120-global-carbon-capture-and-storage-ca
[294] Lackner K.S., Ziock H.-.J. Grimes P. (1999), Carbon Dioxide Extraction from Air: pacity-grew-by-33-in-2020-says-new-report (accessed Dec. 14, 2020).
Is it an Option? Technical Report LA-UR-99–583 (Los Alamos National [325] DeVries T, Holzer M, Primeau F. Recent increase in oceanic carbon uptake driven
Laboratory). by weaker upper-ocean overturning. Nature 2017;542:215–8.
[295] Lackner KS. The promise of negative emissions. Science 2016;354:714. [326] Rau GH, Caldeira K. Enhanced carbonate dissolution: a means of sequestering
[296] Wilcox J, Psarras PC, Liguori S. Assessment of reasonable opportunities for direct waste CO2 as ocean bicarbonate. Energy Convers. Manag. 1999;40:1803–13.
air capture. Environ. Res. Lett. 2017;12:065001. [327] Caldeira K, Rau GH. Accelerating carbonate dissolution to sequester carbon
[297] Sanz-Perez ES, Murdock CR, Didas SA, Jones CW. Direct capture of CO2 from dioxide in the ocean: geochemical implications. Geophys. Res. Lett. 2000;27:
ambient air. Chem. Rev. 2016;116:11840–76. 225–8.
[298] Realmonte G, Drouet L, Gambhir A, Glynn J, Hawkes A, Koberle AC, Tavoni M. [328] U.S. National Academy of Sciences report. Climate intervention: reflecting
An inter-model assessment of the role of direct air capture in deep mitigation sunlight to cool earth. National Academy Press; 2015.
pathways. Nat. Commun. 2019;10:3277. [329] Bronselaer B, Zanna L. Heat and carbon coupling reveals ocean warming due to
[299] Deployment of Deep Decarbonization Technologies. Proceedings of a workshop, circulation changes. Nature 2020;584:227–33.
wash. D.C. The National Academies Press; 2019. http://nap.edu/25656. accessed [330] Bushuyev OS, De Luna P, Dinh CT, Tao L, Saur G, van de Lagemaat J, SO Kelley,
Dec 12, 2020. Sargent EH. What should we make with CO2 and how can we make it? Joule
[300] Keith DW. Why capture CO2 from the atmosphere? Science 2009;325:1654–5. 2018;2:825–32.
[301] Keith DW, Holmes G, St. Angelo D, Heidel K. A process for capturing CO2 from the [331] Haung CH, Tan CS. A review: CO2 utilization. Aerosol Air Qual. Res. 2014;14:
atmosphere. Joule 2018;2:1573–94. 480–99.
[302] Forbes, Dec. 2, 2020, The future of carbon capture is in the air, https://www.fo [332] Graves C, Ebbesen SD, Mogensen M, Lackner KS. Sustainable hydrocarbon fuels
rbes.com/sites/feliciajackson/2020/12/02/is-the-future-of-carbon-capture-in-th by recycling CO2 and H2O with renewable or nuclear energy. Renew. Sustain.
e-air/?sh=20fc319e2c4a (accessed Dec. 5, 2020). Energy Rev. 2011;15:1–23.
[303] Munkejord ST, Hammer M, Lovseyh SW. CO2 transport: data and models – A [333] Kumaravel V, Bartlett J, Pillai SC. Photoelectrochemical conversion of carbon
review. Appl. Energy 2016;169:499–523. dioxide (CO2) into fuels and value-added products. ACS Energy Lett 2020;5:
[304] Roussanaly S, Jakobsen JP, Hognes ES, Brunsvold AL. Benchmarking of CO2 486–519.
transport technologies: part I – Onshore pipeline and shipping between two [334] Tu W, Zhou Y, Zou Z. Photocatalytic conversion of CO2 into renewable
onshore areas. Intern. J. Greenhouse Gas Control 2013;19:584–94. hydrocarbon fuels: state-of-the-art accomplishment, challenges, and prospects.
[305] Roussanaly S, Brunsvold AL, Hognes ES. Benchmarking of CO2 transport Adv. Mater. 2014;26:4607–26.
technologies: part II – Offshore pipeline and shipping to an offshore site. Intern. J. [335] Dieterich V, Buttler A, Hanel A, Spliethoff H, Fendt S. Power-to-liquid via
Greenhouse Gas Control 2014;28:283–99. synthesis of methanol, DME or Fischer-Tropsch fuels: a review. Energy Environ.
[306] Aspelund A, Molnvik MJ, De Koeijer G. Ship transport of CO2: technical solutions Sci. 2020;13:3207–52.
and analysis of costs, energy utilization, exergy efficiency and CO2 emissions. [336] Dipietro P., Improving Domestic Energy Security and Lowering CO2 Emissions
Chem. Eng. Res. Design 2006;84:847–55. with “Improving Domestic Energy Security and Lowering CO2 Emissions with
[307] Al Baroudi H, Awoyomi A, Patchigolla K, Jonnalagadda K, Anthony EJ. A review “Next Generation” CO2-Enhanced Oil Recovery (CO2-EOR)”, Report # DOE/
of large-scale CO2 shipping and marine emissions management for carbon NETL-2011/1504, June 2011, http://www.netl.doe.gov/energy-analyses/pubs/
capture, utilization and storage. Appl. Energy 2021;287. 116510 1-42. NextGen_CO2_EOR_06142011.pdf.
[308] Cole IS, Corrigan P, Sim S, Birbilis N. Corrosion of pipelines used for CO2 [337] Technology Roadmap: Carbon Capture and Storage in Industrial Applications
transport in CCS: is it a real problem? Intern. J. Greenhouse Gas Control 2011;5: (2011), International Energy Agency, and United Nations Industrial Development
749–56. Organization, https://www.iea.org/reports/technology-roadmap-carbon-capture
-and-storage-in-industrial-applications (accessed Jan. 14, 2021).

54
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

[338] DeSimone JM. Practical approaches to green solvents. Science 2002;297: [371] Hori Y, Kikuchi K, Suzuki S. Production of CO and CH4 in electrochemical
799–803. reduction of CO2 at metal electrodes in aqueous hydrogen carbonate solution.
[339] Gotz M, Lefebvre J, Mors F, Koch AM, Graf F, Bajohr S, Reimert R, Kolb T. Chem. Lett. 1985;11:1695–8.
Renewable power-to-gas: a technological and economic review. Renew. Energy [372] Whipple DT, Kenis PJA. Prospects of CO2 utilization via direct heterogeneous
2016;85:1371–90. electrochemical reduction. J. Phys. Chem. Lett. 2010;1:3451–8.
[340] Buttler A, Spliethoff H. Current status of water electrolysis for energy storage, grid [373] Kauffman DR, Thakkar J, Siva R, Matranga C, Ohudnicki PR, Zeng C, Jin R.
balancing and sector coupling via power-to-gas and power-to-liquids: a review. Efficient electrochemical CO2 conversion powered by renewable energy. ACS
Renew. Sustain. Energy Rev. 2018;82:2440–54. Appl. Mater. Interfaces 2015;7:15626–32.
[341] Blanco H, Faaij A. A review at the role of storage in energy systems with a focus [374] Lu Q, Jiao F. Electrochemical CO2 reduction: Electrocatalysts, reaction
on power to gas and long-term storage. Renew. Sustain. Energy Rev. 2018;81: mechanism, and process engineering 2016;29:439–56.
1049–86. [375] Weng Z, Jiang J, Wu Y, Wu Z, Guo X, Materna KL, Liu W, Batista VS, Brudvig GW,
[342] Bailera M, Lisbona P, Romeo LM, Espatolero S. Power to gas projects review: lab, Wang H. Electrochemical CO2 reduction to hydrocarbons on a heterogeneous
pilot, and demo plants for storing renewable energy and CO2. Renew. Sustain. molecular Cu catalyst in aqueous solution. J. Am. Chem. Soc. 2016;138:8076–9.
Energy Rev. 2018;69:292–312. [376] Costentin C, Robert M, Saveant J-M. Catalysis of the electrochemical reduction of
[343] Thema M, Bauer F, Sterner M. Power-to-gas: electrolysis and methanation status carbon dioxide. Chem. Soc. Rev. 2013:2423–36.
review. Renew. Sustain. Energy Rev. 2019;112:775–87. [377] Dubois MR, Dubois DL. Development of molecular electrocatalysts for CO2
[344] Parra D, Zhang X, Bauer C, Patel MK. An integrated techno-economical and life reduction and H2 production/oxidation. Accnts. Chem Res. 2009;42:1974–82.
cycle environment assessment of power-to-gas systems. Appl. Energy 2017;193: [378] Kortlever R, Shen J, Schouten KJP, Calle-Vallejo F, Koper MTM. Catalysis and
440–54. reduction pathways for the electrochemical reduction of carbon dioxide. J. Phys.
[345] De Luna P, Hahn C, Higgins D, Jaffer SA, Jaramillo TF, Sargent EH. What would it Chem. Lett. 2015;6:4073–82.
take for renewably powered electrosynthesis to displace petrochemical processes? [379] Liu C, Cundari TR, Wilson AK. CO2 reduction on transition metal (Fe, Co, Ni, and
Science 2019;364(3506):1–9. Cu) surfaces: in comparison with homogeneous catalysis. J. Phys. Chem. C 2012;
[346] Lee WJ, Li C, Prajitno H, Yoo J, Patel J, Yang Y, Lim S. Recent trend in thermal 116:5681–8.
catalytic low temperature CO2 methanation: a critical review. Catal. Today 2021; [380] Centi G, Perathoner S. Catalysis: role and challenges for a sustainable energy. Top.
368:2–19. Catal. 2009;52:948–61.
[347] Su X, Yang X-F, Huang Y, Liu B, Zhang T. Single-atom catalysis toward efficient [381] NETL CO2U lca guidance toolkit. National Energy Technology Laboratory, U.S.
CO2 conversion to CO and formate products. Acc. Chem. Res. 2019;52:656–64. Department of Energy; 2021. https://www.netl.doe.gov/LCA/CO2U. accessed
[348] Bhosale RR, Takalkar G, Sutar P, Kumar A, AlMomani F, Khraisheh M. A decade of Jan 19.
ceria based solar thermochemical H2O/CO2 splitting cycle. Intern. J. Hydrogen [382] Stiglitz JE, Stern N, Duan M, Edenhofer O, Giraud G, Heal GM, La Rovere EL,
Energy 2019;44:34–60. Morris A, Moyer E, Pangestu M, Shukla PR, Sokona Y, Winkler H. Report of the
[349] Snoeckx R, Bogaerts A. Plasma technology – a novel solution for CO2 conversion. high-level commission on carbon prices. Washington DC: The World Bank; 2017.
Chem. Soc. Rev. 2017;46:5805–63. [383] Burke M, Craxton M, Kolstad CD, Onda C, Allcott H, Baker E, Barrage L, Carson R,
[350] Ashford B, Tu X. Non-thermal plasma technology for the conversion of CO2. Gillingham K, Graff-Zivin J, Greenstone M, Hallegate S, Hanemann WM, Heal G,
Current Opinion Green Sustain. Chem. 2017;3:45–9. Hsiang S, Jones B, Kelly DL, Kopp R, Kotchen M, Mendelsohn R, Meng K,
[351] Li X, Yu J, Jaroniec M, Chen X. Cocatalysts for selective photoreduction of CO2 Metcalf G, Morena-Cruz J, Pindyck R, Rose S, Rudik I, Stock J, Tol RSJ.
into solar fuels. Chem. Rev. 2019;119:3962–4179. Opportunities for advances in climate change economics. Science 2016;352:
[352] Kuehnel MF, Orchard KL, Dalle KE, Reisner E. Selective photocatalytic CO2 292–3.
reduction in water through anchoring of a molecular Ni catalyst of CdS [384] Rosenbloom D, Markard J, Geels FW, Fuenfschilling L. Why carbon pricing is not
nanocrystals. J. Am. Chem. Soc. 2017;139:7217–23. sufficient to mitigate climate change – and how “sustainability transition policy”
[353] Nitopi S, Bertheussen E, Scott SB, Liu X, Engstfeld AK, Horch S, Seger B, can help. Proceed. National Acad. Sci. 2020;117:8664–8.
Stephens IEL, Cham K, Hahn C, Norskov JK, Jaramillo TF, Chorkendorff I. [385] Carbon Pricing Dashboard (Nov. 1, 2020), The World Bank, https://carbonpricing
Progress and perspectives of electrochemical CO2 reduction on copper in aqueous dashboard.worldbank.org/map_data (accessed Jan 19, 2021).
electrolytes. Chem. Rev. 2019;119:7610–72. [386] Kolstad CD. Who pays for climate regulation? Stanford Institute for Economic
[354] Kungas R. Review – Electrochemical CO2 reduction for CO production: Policy Research ((SIEPR); 2014. SIEPR Policy BriefJan. 2014, https://siepr.st
comparison of low- and high-tempearture technologies. J. Electrochem. Soc. anford.edu/sites/default/files/publications/SIEPR_PolicyBrief_Kolstad_v4_2.pdf.
2020;167:044508. accessed Jan. 15, 2021.
[355] Zheng Y, Wang J, Yu B, Zhang W, Chen J, Qiao J, Zhang J. A review of high [387] Report of the High-Level Commission on Carbon Prices, May 29, 2017, Carbon
temperature co-electrolysis of H2O and CO2 to produce sustainable fuels using Pricing Coalition Leadership, The World Bank, https://www.carbonpricingleade
solid oxide electrolysis cells (SOECs): advanced materials and technology. Chem. rship.org/report-of-the-highlevel-commission-on-carbon-prices (accessed April 9,
Soc. Rev. 2017;46:1427–63. 2021).
[356] Motoya JH, Seitz LC, Chakthranont P, Vojvodic A, Jaramillo TF, Norskov JK. [388] Why carbon pricing isn’t working?, Foreign Affairs, July/August 2018, https
Materials for solar fuel and chemicals. Nature Mater 2017;16:70–81. ://www.foreignaffairs.com/articles/world/2018-06-14/why-carbon-pricing-isn
[357] Kumar B, Liorente M, Froehlich J, Dag T, Sathrum A, Kubiak CP. Photochemical t-working (accessed Jan 15, 2021).
and photoelectrochemical reduction of CO2. Ann. Rev. Phys. Chem. 2012;63: [389] IPCC 2011 Special Report of renewable energy sources and climate change
541–69. mitigation (SRREN), June 28, 2011, https://www.ipcc.ch/2011/06/28/special-re
[358] Xu S, Carter EA. Theoretical insights into heterogeneous (photo)electrochemical port-on-renewable-energy-sources-and-climate-change-mitigation-srren/
CO2 reduction. Chem. Rev. 2019;119:6631–69. (accessed Dec 20, 2020).
[359] Dry ME. The Fischer-Tropsch process: 1950-2000. Catal. Today 2002;71:227–41. [390] Larcher D, Tarascon J-M. Towards greener and more sustainable batteries for
[360] Rofer-DePoorter CK. A comprehensive mechanism for the Fischer-Tropsch electrical energy storage. Nature Chem 2015;7:19–29.
synthesis. Chem. Rev. 1981;81:447–74. [391] IPCC 2011, Special Report on renewable energy sources and climate change
[361] Jahangiri H, Bennett J, Mahjoubi P, Wilson K, Gu S. A review of advanced catalyst mitigation (SRREN) June 28, 2011 https://www.ipcc.ch/site/assets/uploads/20
development for Fischer-Tropsch synthesis of hydrocarbons from biomass derived 18/03/SRREN_Full_Report-1.pdf (accessed Feb 13, 2021).
syn-gas. Catal. Sci. Technol. 2014;4:2210–29. [392] Carbajales-Dale M, Barnhart CJ, Brandt AR, Benson SM. A better currency for
[362] Ni Y, Chen Z, Fu Y, Liu Y, Zhu W, Liu Z. Selective conversion of CO2 and H2 into investing in a sustainable future. Nat Clim Chang 2014;4:524–7.
aromatics. Nature Commun 2018;9(3457):1–7. [393] Barnhart CJ, Dale M, Brandt AR, Benson SM. The energetic implications of
[363] Gür TM, Huggins RA. Methane synthesis on nickel by a solid state ionic method. curtailing versus storing solar- and wind-generated electricity. Energy Environ.
Science 1983;219:967–9. Sci. 2013;6:2804–10.
[364] Gür TM, Huggins RA. Methane synthesis over transition metal electrodes in a [394] Carbajales-Dale M, Barnhart CJ, Benson SM. Can we afford storage? A dynamic
solid state ionic cell. J. Catal. 1986;102:443–6. net energy analysis of renewable electricity generation supported by energy
[365] Gür TM, Wise H, Huggins RA. Electrocatalytic conversion of carbon dioxide to storage. Energy Environ. Sci. 2014;7:1538–44.
methane and oxygen with an oxide ion-conducting electrolyte. J. Catal. 1991;129: [395] The Emissions Gap Report 2014: A UNEP Synthesis Report, http://www.unep.
216–24. org/publications/ebooks/emissionsgapreport2014/.
[366] Gür TM, Nohmi T, Wise H, Huggins RA. Recovery of Oxygen from Carbon Dioxide [396] IPCC-AR5 report 2014 (Mitigation of Climate Change - 5th Assessment Report:
in a High-Temperature Electrochemical Reactor. J. Electrochem. Soc. 1988;135: chapter 7. Energy Systems), https://www.ipcc.ch/report/ar5/wg3/,
C343. https://www.ipcc.ch/pdf/assessment-report/ar5/wg3/ipcc_wg3_ar5_chapter7.
[367] Tao G, Sridhar KR, Chan CL. Study of carbon dioxide electrolysis at electrode/ pdf.
electrolyte interface: part I. Pt/YSZ interface. Solid State Ionics 2004;175:615–9. [397] Coren MJ. Solar and wind are now the cheapest energy around – unless you need
[368] Tao G, Sridhar KR, Chan CL. Study of carbon dioxide electrolysis at electrode/ to store it. Quartz 2017. Nov. 14, https://qz.com/1125355/solar-and-wind-are-
electrolyte interface: part II. Pt-YSZ cermet/YSZ interface. Solid State Ionics 2004; now-the-cheapest-energy-around-unless-you-need-to-store-it/. accessed June 3,
175:621–4. 2021.
[369] Alemozafar AR, Gür TM, Homsy GM. Solid State Electrochemical Oxygen [398] Wansted D, Schlissel D. Petra-Nova mothballing post-mortem: closure of texas
Conversion For Martian And Lunar Environments. In: Proceedings of the NASA carbon capture plant is a warning sign. Institute for Energy Economics and
Microgravity Materials Science Conference; 2000. June 6-8, 2000. Financial Analysis; Aug. 2020. https://ieefa.org/wp-content/uploads
[370] Sridhar KR, Iacomini CS. Combined H2O/CO2 solid oxide electrolysis for Mars in /2020/08/Petra-Nova-Mothballing-Post-Mortem_August-2020.pdf (accessed
situ resource utilization. J. Propulsion Power 2004;20:892. June 5, 2021.

55
T.M. Gür Progress in Energy and Combustion Science 89 (2022) 100965

[399] 2030 or Bust: 5 key takeaways from the IPCC report (Oct. 18, 2018), The Climate Turgut M. Gür is an Adjunct Professor of Materials Science and
Reality Project, https://www.climaterealityproject.org/blog/2030-or-bus Engineering at Stanford University, and a recognized leader in
t-5-key-takeaways-ipcc-report. high temperature electrochemical energy conversion and
[400] US National Oceanic and Atmospheric Administration (NOAA), https://www.cli storage materials and technologies with 11 US issued patents
mate.gov/print/833067 (accessed Dec 14, 2020). and 160 published articles. He is currently the Senior Vice
[401] US-DOE Office of Science, Basic Energy Sciences, https://science.osti.gov/ President, Board Director, and Fellow of The Electrochemical
-/media/budget/pdf/sc-budget-request-to-congress/fy-2022/FY_2022_SC_BES_ Society. He holds BS and MS degrees in chemical engineering
Cong_Budget.pdf?la=en&hash=FBE9A9F26A629F9E2F0C42 from Middle East Technical University (METU) in Ankara/
55FB48D8E04A4B44B8 (accessed June 6, 2021). Turkey, and a Ph.D. degree in materials science and engi­
[402] National Academies of Sciences, Engineering, and Medicine. In: Deployment of neering from Stanford University.
Deep Decarbonization Technologies: Proceedings of a Workshop. Washington,
DC: The National Academies Press; 2019. https://doi.org/10.17226/2565or,
http://nap.edu/25656. accessed Dec. 11, 2020.

56

You might also like