You are on page 1of 11

Applied Energy 325 (2022) 119895

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

The impact of climate on solvent-based direct air capture systems


Keju An a, Azharuddin Farooqui b, Sean T. McCoy b, *
a
Buildings and Transportation Science Division, Oak Ridge National Laboratory, Oak Ridge, TN 37830, USA
b
Department of Chemical and Petroleum Engineering, University of Calgary, AB T2N IN4, Canada

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• The CO2 capture rate of liquid-solvent


DAC is strongly influenced by air tem­
perature and relative humidity.
• Varying CO2 capture rates have a strong
impact on overall energy demand of
DAC and levelized cost of CO2 removal.
• Levelized cost of capture can vary by a
factor of nearly two between cold, dry
conditions and warm, humid conditions.
• GHG intensity of energy supply is,
however, more influential than climate
conditions in determining efficiency of
removal.
• Estimates of DAC potential should
consider specific climate conditions
rather than relying on generic perfor­
mance figures.

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

Keywords: Direct air capture (DAC) is increasingly seen as a critical technology to reach mid-century net-zero targets and
Direct Air Capture (DAC) limit climate change to well below 2 ◦ C. While the commercialization of DAC technologies is being pursued by
Climate impact numerous companies, there remains remarkably little information on their performance under “real-world”
CO2 capture efficiency
conditions. In this paper, for the first time, we investigate the influence of temperature and relative humidity of
Levelized cost for CO2 capture
air on the CO2 capture rate at the air contactor, overall energy requirement, CO2 capture efficiency, and levelized
cost of liquid-solvent based DAC systems. We observe that the overall energy demand decreases from 11.1 to 8.3
GJ/tCO2 as the CO2 capture rate increases from 40 to 85 % and that high capture rates can only be achieved in
hot and humid climate conditions. We observe that a CO2 capture rate of 75 % is only possible above 17 ◦ C and
90 % relative humidity, and this drops dramatically at lower temperatures. It is also observed that water
evaporation in the air contactor is highest at dry and low relative humidity, as expected. The sensitivity analysis
showed that CO2 capture efficiency is relatively insensitive to climate conditions for the liquid-solvent based DAC
plant. Lastly, the levelized cost of natural gas standalone scenario varies from $240/tCO2 to $409/tCO2, and this
is more sensitive to temperature than relative humidity. The levelized cost of the natural gas stand-alone case is
between 7 % and 10 % lower than that of an electric grid-connected case across all climate conditions.

* Corresponding author.
E-mail address: sean.mccoy@ucalgary.ca (S.T. McCoy).

https://doi.org/10.1016/j.apenergy.2022.119895
Received 1 April 2022; Received in revised form 16 August 2022; Accepted 24 August 2022
Available online 6 September 2022
0306-2619/© 2022 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
K. An et al. Applied Energy 325 (2022) 119895

1. Introduction echoed by Sabatino et al. [23], who used process modeling to analyze
solid-sorbent and liquid-sorbent DAC. They found that despite the
Greenhouse gas (GHG) emissions from human activities are driving higher exergy demand and productivity for solid sorbents, both have the
unprecedented change in the climate system that is manifesting itself in potential to achieve costs below $200/tCO2. In addition to the above
rapid, widespread warming [1] and extreme events like wildfires, heat works, the economics of DAC and policy levers necessary to drive
waves, heavy rain, and droughts [2]. While many greenhouse gases, deployment and drive cost reductions through “learning-by-doing” are
such as methane and nitrous oxide, contribute to radiative forcing, being increasingly investigated [7,24–28].
carbon dioxide (CO2) is the main contributor. Most climate stabilization Most of the studies summarized above investigated DAC under
scenarios that limit global warming to well below 2 ◦ C include negative relatively moderate climate conditions (e.g., 15–27 ◦ C and 69 % relative
emission technologies (NETs), which remove and isolate CO2 from the humidity [10,17,29]) or provide no specific climate information about
atmosphere [3]. An analysis of scenarios developed using integrated absorption/adsorption processes [9,30,31]. Climate conditions are one
assessment models (IAMs) suggests that a more stringent 1.5 ◦ C warming of the key variables for a DAC system since, unlike with other forms of
limit is unlikely to be feasible without the large-scale deployment of CO2 capture, the CO2 source is the atmosphere that, globally, presents a
NETs and, while some scenarios limit warming to 2.0 ◦ C without NETs, wide range of temperature and humidity. Previous studies of absorption
these require immediate and strong emissions reductions [3–5]. Evi­ kinetics in the CO2 contactor [23,32,33] showed a strong temperature
dence also suggests a portfolio of NETs may be preferable, in particular, dependence. The large volume of air that is contacting the solvent in the
to manage the risks that come with large-scale deployment [3,6,7]. air contactor will cause the solvent to rapidly equilibrate with the air
Direct air capture (DAC) is an engineered system that captures car­ temperature. Parameters such as wind speed, precipitation, and altitude
bon dioxide directly from the atmosphere and could play a prominent will also play a role in determining the performance of DAC, but tem­
role in the NET portfolio [6]. DAC has been increasingly discussed as a perature and relative humidity (RH) should have a significant effect on
climate change mitigation option because of its relatively low land-use kinetics in the air contactor, which has a strong influence on the carbon
intensity and the flexibility in siting DAC. Because the CO2 concentra­ capture rates and overall energy demand on the process. None of the
tion in the atmosphere (about 415 ppm in 2021) is more or less the same peer-reviewed studies published to date explicitly address the impact of
in any location around the planet, DAC could be located close to geologic air temperature and relative humidity (RH) on the system performance
CO2 storage options and energy sources [6,8]. However, estimates of of DAC. By examining the performance of DAC under various climate
current (and future) DAC system energy demand, capital andoperation conditions, spatially resolved estimates of CO2 capture rates, CO2 cap­
cost, and water consumption vary widely [9–11]. There are a few ture efficiencies, and cost could be developed that could influence the
companies (e.g., Climeworks, Carbon Engineering, Global Thermostat) deployment of liquid-solvent based DAC systems—and, possibly, shift
commercially developing different DAC technologies. Solid sorbents the preference between different types of DAC systems.
with low-temperature regeneration (~120 ◦ C) [8,11–17] and liquid Liquid DAC systems require both high-temperature heat and elec­
solvents with high-temperature regeneration (~900 ◦ C) [18–21] are tricity, with heat accounting for over 80 % of the energy demand. In
furthest in development and provide a scalable opportunity to evaluate systems described in the literature, electrical demand has been met
DAC technology. Adsorption-based systems can employ a wide range of through co-generation (e.g., using a natural gas combined cycle unit) or
sorbents that are tailored to the application and, because of the low- import of electricity from the grid, and the thermal demand met largely
temperature heat requirement, can use a wide range of energy sour­ by oxyfuel combustion of natural gas. McQueen et al. [19] and Liu et al.
ces—including waste heat, geothermal energy, and renewable elec­ [31] investigated the influence of providing heat using electricity
tricity [9,19,22]. A challenge for solid sorbent systems is to build a very generated by a range of different technologies, (e.g., geothermal, solar,
large surface area structure with low-pressure drop at a low cost that can wind, nuclear) to reduce the overall life cycle GHG emissions for fixed
also be efficiently heated and sealed (to allow for temperature-vacuum temperature and relative humidity conditions. However, these studies
swing regeneration). Additionally, there are trade-offs in sorbent do not identify whether the life cycle GHG emissions and levelized cost
design between scalability, cost, and lifetime over numerous adsorp­ of liquid DAC systems are more sensitive to the fuel (including fugitive
tion–desorption cycles. In contrast, liquid solvent DAC allows for emissions) or the climate conditions (i.e., temperature and relative
continuous operation and uses relatively conventional cooling-tower humidity).
systems for air contacting. The bottleneck of liquid solvent DAC is the There have also been several recent life cycle assessment (LCA)
high energy demand for the calcining process used in solvent regener­ studies to quantify the environmental performance of DAC. Deutz et al.
ation, the relative complexity of the system, and relatively large water [17] showed the current DAC plants in Hellisheidi (Iceland), and Hinwii
consumption in dry climates [10]. To apply the DAC system at large (Switzerland), operated by Climeworks (based on solid sorbents) can
commercial scales, low-cost and efficient systems are essential. already deliver negative emissions with a carbon capture efficiency of
Multiple papers published in recent years present the results of 85.4% and 93.1%, respectively. In addition, they show that a low-GHG
modeling and demonstrations of liquid-based DAC systems. Holmes intensity energy source and high efficiency are essential for lowering the
et al. [18] presented the results of a novel air contactor utilizing a life cycle carbon footprint of the solid sorbent-based DAC system. Ter­
crossflow configuration and high-performance packing material. This louw et al. [30] presented a comprehensive LCA of different solid-
prototype air contactor significantly reduced the fan power required to sorbent DAC systems with low-carbon electricity and heat sources.
push ambient air through the air contactor as well as lower the pressure The sensitivity analysis reveals the importance of selecting appropriate
drop, which is closely related to the operational cost. Keith et al. [10] regions because of the carbon intensity of grid electricity that can lead to
reported the energy and material balance of a commercial scale (1 Mt/ a net positive emission system today. Madhu et al. [29] presented a
year) liquid-based DAC system based on process simulation and pilot comparative life cycle assessment of both DAC technologies coupled
plant results, including all of the major components (air contacting, with carbon storage: liquid-solvent and solid sorbent systems. Their
solvent regeneration, and CO2 compression). In addition, they estimated results suggest that solid-sorbent DAC outperforms liquid-solvent DAC
energy demand (5.53 GJ/tCO2 to 8.81 GJ/tCO2) and levelized capture by a factor of 1.3–10 in all environmental impact scenarios for the 1Mt-
cost ($94/tCO2 to $232/tCO2) for different energy supply options and CO2/year scale. However, for the liquid-solvent DAC, Madhu et al. [29]
economic parameters. The experimental and pilot plant results pub­ use a capture rate, of 42% across from ambient air which is a far lower
lished by Holmes et al. [18] and Keith et al. [10] form the basis for much capture rate than in Keith et al. [10] across all cases, which is sub-
of the analysis of liquid-solvent DAC systems. Their results show that the optimal based on Holmes’s work [18,20] (suggesting 75 %). In addi­
estimated energy demand is relatively high but that sensible costs might tion, Madhu et al. [29] do not capture CO2 emissions associated with the
be achievable with further technology development. These findings are energy used in calcination, which implies a different process than

2
K. An et al. Applied Energy 325 (2022) 119895

previously proposed for liquid-solvent DAC systems. Liu et al. [31] Engineering process [10]. The mass and energy balances at different CO2
conducted an LCA of the greenhouse gas emissions of a liquid-solvent capture rates and the thermodynamic transformations of the operating
DAC system paired with Fischer–Tropsch synthesis (FTS) to produce streams were obtained using the commercial process simulator Aspen
transportation fuel. This work was the first LCA study of a DAC-to-fuel Plus V11. The assumptions considered for each stream and component
process and the system emits 0.51 gCO2e per gCO2 captured from the and the process parameters adopted to derive reasonable results are
air but this result is extremely sensitive to upstream electricity emis­ provided in the following sections. Fig. 1 shows the process flow dia­
sions. Previous LCA studies have shown the strong influence of energy gram of the DAC plant with an electrical support option of an NGCC unit
source and capture efficiency on the net of lifecycle emissions associated or grid electricity.
with DAC systems. The growing literature on overall LCA on DAC system
either talk about the alternative energy source for supporting the energy 2.1. Thermodynamic model and chemical reactions
demand to reduce the GHG emissions or levelized the cost of CO2 cap­
ture without considering the influence of varying temperatures and The thermodynamic parameters and properties of the chemicals
relative humidity. were estimated with the ENRTL-RK package for the liquid phase and RK-
The present study addresses this gap and investigates (1) the SOAVE and Henry’s Law for the gas phase [34–36]. The unsymmetrical
response of CO2 capture rates in a solvent-based DAC system to a range electrolyte NRTL property method (ENRTL-RK) is utilized for the binary
of temperature and humidity conditions and the associated change in interaction parameters, electrolyte pair parameters, and equilibrium
the levelized cost of CO2 capture, energy demand, and net CO2 capture constant for the precipitation reaction estimated from experimental data
rates and (2) the impact of meeting electrical demand from the grid [37]. For the two-phase blocks, the mass transfer was modelled with a
rather than a standalone natural gas combined cycle (NGCC) unit. To two-film rate-based approach neglecting the mixing by convection [21].
achieve these objectives, we model a solvent-based DAC system­ This thermodynamic approach estimates the excess Gibbs free energy
—including the air contactor, regeneration system, and CO2 compres­ and the activity coefficient of an electrolyte system, considering the
sion, validated against Carbon Engineering data [10]—and use this strong non-ideal behaviour of ionic mixtures. The excess of Gibbs energy
model to develop mass and energy balances across a range of air tem­ can be defined instead as a difference between the Gibbs free energy of
perature and relative humidity to estimate the CO2 capture efficiency. electrolyte systems and the Gibbs energy of an ideal solution with the
Since the net CO2 capture from the atmosphere (across the life cycle) is same condition of temperature, pressure, and concentration [38].
of critical importance for DAC, we include the influence of natural gas Chemical reactions in each component involve not only equilibrium
upstream fugitive emissions in our estimates. In the grid-connected case, reactions but also the salt formation and dissociation reactions as shown
we remove the NGCC unit and use imported grid electricity to meet all in Table 1. The series of reactions have been generated by the Aspen Plus
electrical energy demand from the system as shown in Fig. 1. These with the procedure of electrolyte wizard that all the chemical substances
results will contribute to the understanding of DAC under different present in the DAC system including the ionic components and its re­
climate conditions and the necessary conditions for DAC to deliver the actions with coefficients to calculate equilibrium constants.
intended climate benefit to operate as net negative carbon capture
technology. 2.2. Process layout

2. Materials and methods The DAC process was modelled in two separate loops: first, CO2
capture and regeneration cycle using an aqueous solution and, second, a
The process simulation of the liquid solvent DAC system includes the calcium loop drives the removal of carbonate ion. The liquid solvent-
air contacting unit, regeneration, and calcium cycles to capture CO2 based DAC plant captures roughly 1 Mt of CO2 in a year from the at­
from ambient air in a continuous mode and is based on the Carbon mosphere and delivers dry CO2 at 15 MPa after four stages of

Fig. 1. Commercial-scale, 1 MtCO2/year, DAC plant configuration [10] illustrated as a block flow diagram showing the option of electricity supply by the NGCC unit
(inside the dashed box) or grid electricity.

3
K. An et al. Applied Energy 325 (2022) 119895

Table 1 Table 2
Chemical reactions in the DAC system. Operating parameters and specifications for rated-based air-contactor simula­
Reaction Type Reaction Type
tion [10,23].
Operating Condition of rate based Air-contactor
(1)H2 O + Equilibrium (5)K2 CO3 ↔ CO2−
3 + 2K
+ Salt
HCO−3 ↔ CO2−
3 +
Precipitation Air flow rate [tonne/h] 150.6
H 3 O+ Lean solvent flow rate [tonne/h] 19.08
(2)CaOH+ ↔ Ca2+ + Equilibrium (6)CaCO3 ↔ CO2−
3 +
Salt KOH concentration in the lean solvent [mol/Liter] 1.1
OH− Ca2+ Precipitation K2CO3 concentration in the lean solvent [mol/Liter] 0.45
(3)2H2 O + Equilibrium (7)KOH ↔ OH− + K+ Dissociation Packing Diameter [m] 5.64
CO2 ↔ HCO−3 + Packing Depth [m] 1.28
H 3 O+ Packing Type Sulzer 250Y
(4)2H2 O ↔ OH− + Equilibrium (8)Ca(OH)2 ↔ CaOH+ + Dissociation Number of blocks 6
H 3 O+ OH− Reaction condition factor 0.9
Film discretization points 5
Interfacial area factor 1.2
Liquid film discretization points 5
compression. The air contactor design is based on the commercially
CO2 Capture Rate [%] 76.9
available crossflow cooling tower and this design choice is critical for
cost-effective DAC on a commercial scale [10]. The air-contactor leads
ambient air in contact with the liquid solution and captures CO2 from solvent are designed based on a scaling factor with the lean solvent
the air. As the liquid solvent flows down the packing column, it coats the molarity of 1.1 mol/L, KOH, and 0.45 mol/L, K2CO3. The modeled
material with a thin film (~50 µm) that allows for the mass transfer of performance of CO2 capture in the air contactor was found to be 76.9%
CO2 from the gas to the liquid phase. It follows that the degree of CO2 compared to 75% from Keith et al. [10] for the same conditions (i.e., a
separation from the gas mixture increases with the larger packing vol­ 2.5% difference). We then simulate performance of theair contactor
ume [21] After the CO2 is captured in a dedicated air contactor unit, under a wide range of temperature and humidity climate conditions
where it is absorbed in an aqueous solution of KOH in the form of K2CO3, could affect CO2 capture rates notably, based on the previous works of
a rich solvent. The K2CO3 solution is regenerated by forming a calcium absorption kinetics [32]. The detailed methodology of the model vali­
carbonate (CaCO3) which is then fed to an oxygen-fired calciner dation of the air contactor and the chemical reactions are listed in the
(~900 ◦ C) and decomposed to CaO and CO2. Carbon dioxide captured supplementary file (Section 1 and Table S2).
from both the atmosphere and from fuel oxidization exits the calciner.
2.2.2. DAC system model
2.2.1. Air contactor model The commercial-scale DAC plant (1 MtCO2/year) was modeled in
In the air contactor, ambient air passes over a solvent containing a Aspen Plus to assess the response of energy demand and CO2 capture
hydroxide ion that reacts with the CO2, forming a carbonate compound. efficiency at different CO2 capture rates from the air contactor. Our DAC
The absorption mechanism of CO2 in alkaline solutions is well known model is representative of a commerical configuration [10], but required
and takes place in three main steps, shown by Equations (1)–(3) [32,39] some modifications based on mass, energy, and charge balance to reduce
and other equations that may occur in air contactor are listed in the the model complexity. The main unit components and their operating
supplementary file. conditions are listed in Table 3. For the DAC model, we represent the air-
contactor as a mixer that conceptually reacts the lean solvent with CO2
CO2 + OH − ↔ HCO−3 (1)
in the air to acheive a CO2 capture rate from 40% to 85%. The pellet
reactor was simulated by a crystallizer reactor to remove the carbonate
HCO−3 + OH − ↔ CO2−3 + H2 O (2)

2H2 O ↔ H3 O+ + OH − (3)
Table 3
In this manner, CO2 is captured in the air contactor unit where it is Unit Components and Operating Conditions.
absorbed in an aqueous solution of KOH in the form of K2CO3. During Air Contactor Calciner
the CO2 absorption process, the rate of the reaction between carbon
Block Option Contactor RStoic
dioxide and hydroxide ion increase exponentially as reaction tempera­ Property Method ENRTL-RK PENG-ROB
ture increases at a different level of molar concentrations [21,33]. In the Operating 21 ◦ C 900 ◦ C
air contactor, the ultimate purpose is to maximize the coverage of the Temperature
liquid solvent over the packing column to increase the interfacial contact Operating Pressure 1 bar 1 bar
Reactions (Main reactions) CH4 + O2 → CO2 +
area for optimal gas–liquid reaction. Maximizing the percentage of the
2H2O
wetted packing will provide higher opportunities for CO2 to cross the CO2 + OH– → HCO–3 CaCO3 → CaO + CO2
gas–liquid film interface at which CO2 transfers from the gas to the HCO–3 + OH– → CO–3 + H2O 2C2H6 + 7O2 → 4CO2
liquid phase. In addition, the reaction temperature also affects Henry’s + 6H2O
2H2O → H3O+ + OH– C3H8 + 5O2 → 3CO2 +
constant [40] and diffusion coefficient [41] significantly which are
4H2O
major variables of the mass transfer coefficient formula [20]. In this (see supplementary file for detailed
work, the air contactor has been designed and simulated using Aspen reaction list)
software to validate the CO2 capture rate, 75%, of the Carbon Engi­ Additional Note – Conversion ratio: 98 %
neering system [10] following the approach presented by Sabatino et al. CaCO3
Pellet Reactor Steam Slaker
[23]. The CO2 capture rate can be calculated based on the CO2 mass flow
Block Option Crystallizer RStoic
rate difference between upstream and downstream of air contactor, Property Method ENRTL-RK PENG-ROB
CO2 capture rate = (m ˙
CO2_in− ṁCO2_out )/ṁCO2_in , where mCO2_in is the Operating 21 ◦ C 300 ◦ C
Temperature
mass flowrate of CO2 at the air inlet from the air contactor and mCO2_out
Operating Pressure 1 bar 1 bar
is the outlet mass flowrate of CO2. For example, the CO2 capture rate is Reactions Salt precipitation reaction: Ca+ + CaO + H2O → CaOH2
75% when the upstream CO2 mass flowrate is 400 t/hr and 100 t/hr for CO2–-
3 → CaCO3
the downstream. The specifications of the contactor model are shown in Additional Note – CaO conversion ratio:
85 %
Table 2. In the air-contactor simulation, the flowrate of air and lean

4
K. An et al. Applied Energy 325 (2022) 119895

ion, CO2–3 , formed in the air contactor. The precipitation of calcium also increases. For example, air at RH of 90% at 10 ◦ C the air can take an
carbonate pellets is adapted considering the conversion of the fed Ca additional 0.78 gmoisture/kgair before it is saturated (humidity ratio: 7.66
(OH)2 of roughly 90 %. In the real plant, this solution is recovered from gmoisture/kgair) whereas at 20 ◦ C, the same capacity increases by 1.33
the pellet reactor in a closed loop. To reduce complexity and avoid any gmoisture/kgair to reach saturation state (humidity ratio: 14.75 gmoisture/
convergence problems, the streams containing the solution have been kgair). For 30 ◦ C, the same capacity of air at RH at 90% RH increases by
modelled with an open loop that leads to ease in managing the design 2.84 gmoisture/kgair (from 24.47 gmoisture/kgair to 27.31 gmoisture/kgair) for
specifications of this section. The calcium carbonate pellets are washed, saturation. Figure S2 (b) gives the additional water that air can absorb
dried, pre-heated, and calcined at 900 ◦ C with a CaCO3 to CaO con­ from KOH solvent given the opportunity and, at lower RH and higher
version ratio of 98 %. The energy demand of the calcining process is temperature, it is highest and falls as the RH increases. This leads to
calculated from the natural gas feed using oxygen as the oxidizing agent decrease in water concentration in solvent and thereby increasing KOH
required to reach the adiabatic condition with the target conversion molarity and hence the higher CO2 capture rates. If the reaction rates of
ratio. We use a representative natural gas composition rather than pure KOH with CO2 are only dependent on KOH molarity, then CO2 capture
methane for feeding the calciner and the power island. From the calciner rates (t/h or %) should follow the same trends as that of the evapora­
outlet, the concentrated stream of CO2 is washed in a water knockout tion/additional moisture carrying capacity of air (Figure S2 (b)) but the
and then compressed to 151 bar using multistage compressors. reaction kinetics are more strongly influenced by temperature, meaning
that at a higher temperature the CO2 capture rates would be higher even
3. Results and discussion at higher RH which is reflected from Fig. 2. The water present in solvent
at higher RH and higher temperature leads to water dissociation to form
3.1. Variation in CO2 capture rate at different climate conditions carbonate and bicarbonate as per Equation (3). Therefore, with lesser
evaporation at higher RH and temperature the concentration of water
In this section, the results of the rate-based air contactor from the increases and will dissociate to form more OH– that eventually increases
Aspen simulation provide the various CO2 capture rates under different the molarity and thereby increases the CO2 capture rates.
climate conditions. We simulate the performance of the air-contactor The practical impact of the temperature is that, ceteris paribus, a
system using rate-based simulation between under 0 to 40 ℃ tempera­ DAC plant in cooler and dryer climate conditions would capture less
ture and 10% to 90% relative humidity climate conditions. The tem­ CO2. For example, the average annual temperature was approximately
perature range, 0 to 40 ℃, is determined by the temperature range of 4 ◦ C at 63% RH in Calgary, Canada from March 2020 to March 2021.
published OH– and CO2 kinetic data [32]. Fig. 2 shows the results of the Thus, the air contactor of a DAC system in Calgary would—over the
CO2 capture rate in the air contactor under different climate conditions. course of a year (and under current climate conditions)—remove around
Keith et al. reported a capture rate of 75% for the Carbon Engineering 58% of the CO2 in the air moving through the contactor. Thus, a nominal
system at 21 ◦ C at 69% RH [10] however, across the modeled conditions, 1 MtCO2/y solvent-based DAC plant would remove approximately 0.77
we find that CO2 capture rates could be expected to vary between 52% MtCO2/y. It is not obvious that this could be countered through design
and 90%. The CO2 capture rate is more sensitive to temperature rather measures, such as solvent heating, given the enormous volume of cold
than the relative humidity of air conditions due to the well known air moving through the contactor. In addition to the impact of colder
exponential relationship between temperature and chemical reaction average temperatures, daily temperature extremes can also be highly
rates [32]. Two factors define the effect of temperature and RH. Firstly, variable. For example, in Calgary, Canada, these range from well below
for any given temperature, the water content in the air and the solvent –32 ◦ C to over 25 ◦ C with relative humidity (RH) from 25% to 94%. The
leaving the contactor will approach equilibrium, which is defined by the liquid solvent should remain in a liquid phase down to zero Celsius
humidity ratio (the ability of air to absorb moisture at a certain tem­ (when the mass fraction of KOH is about 1% [10]), which suggests the
perature) and, RH. This will typically (but nor always) result in evapo­ air contactors could operate below 0 ◦ C with a higher mass fraction of
ration of water from the solvent. Secondly, the reaction between KOH [42]. Since the kinetic data is only available down to 0 ◦ C, the
incoming air and KOH produces water (see Equation (2)) as a by- model is unable to predict performance below 0 ◦ C, but we expect the
product. Figure S2 (a and b) shows that with an increase of tempera­ capture rate would quickly fall below 50%.
ture for a specific relative humidity (for example 90%), the humidity
ratio is higher, and the amount of moisture required to reach saturation 3.2. Mass balance and energy demand

The DAC plant discharges approximately one percent of the circu­


lating calcium each cycle as waste in the form of CaCO3. This discharge
flow performs as a purge that controls the buildup of unnecessary ele­
ments that are possibly introduced to the calcium cycle by various routes
[10]. The overall mass flow for each major component is shared in
Table 4. The sources of carbon dioxide in the DAC system are the
ambient air and oxyfuel combustion from calcining process supported
by the natural gas and oxygen from the air separation unit. The natural
gas flow rate to calciner increases as the CO2 capture rate increases to
maintain the adiabatic condition. After the four-stage compression, the
final outflow CO2 rates vary from 104.8 to 189.9 tonne/hr at 40% and
85% CO2 capture rates. The sharp increment of CaCO3 from the pellet
reactor is mainly due to additional incoming Ca2+ ions from the steam
slaker considered by the salt precipitation process of CaCO3 that is
calculated based on the charge balance of OH– and CO2– 3 in the pellet
reactor at different CO2 capture rates.
The energy balances of electric and thermal power consumed at
different CO2 capture rates in the DAC system are provided in Table 5.
Fig. 2. CO2 capture rate at different relative humidity and a temperature range The compressors are the most electricity-intensive component, ac­
from the Air-Contactor unit of the liquid solvent-based DAC system. (Grey- counting for more than 70 % of the total power demand regardless of the
colored area represents unreachable air conditions). CO2 capture rate. The compressors in the DAC plant are utilized for the

5
K. An et al. Applied Energy 325 (2022) 119895

Table 4
Mass balance of major units of DAC at different, climate-dependent CO2 capture rates (Stream numbers are shown in Fig. 1.).
Mass flow rate [tonne/hr] CO2 Capture Rate

50 % 55 % 60 % 65 % 70 % 75 % 80 % 85 %

Air Contactor
1. CO2 inlet 93.2 101.3 109.4 117.5 125.6 133.1 141.8 149.9
2. KOH solvent inlet 35,000 35,000 35,000 35,000 35,000 35,000 35,000 35,000
3. K2CO3 solvent outlet 35,093 35,101 35,109 35,117 35,125 35,133 35,141 35,149
Pellet Reactor
4. K2CO3 solvent inlet 35,093 35,101 35,109 35,117 35,125 35,133 35,141 35,149
5. Ca(OH)2 solvent inlet 710 722 735 748 760 773 786 798
6. CaCO3 makeup 2.5 2.7 2.8 3 3.2 3.4 3.6 3.8
7. CaCO3 seed inlet 4.5 4.9 5.3 5.6 6 6.3 6.7 7
8. CaCO3 disposal 2.5 2.7 2.8 3 3.2 3.4 3.6 3.8
9. CaCO3 outlet 229.4 246.9 264.4 281.8 299.9 316.7 334.2 351.7
KOH solvent inlet 354,000 354,000 354,000 354,000 354,000 354,000 354,000 354,000
KOH solvent outlet 389,000 389,000 389,000 389,000 389,000 389,000 389,000 389,000
Calciner
10. CaCO3 inlet 229.4 246.9 264.4 281.8 299.9 316.7 334.2 351.7
11. O2 inlet 41.7 44.4 48.3 51.5 54.3 57.4 60.6 64.2
12. Natural Gas inlet 9.15 9.75 10.6 11.3 11.9 12.6 13.3 14.1
13. CaCO3 seed outlet 4.5 4.9 5.3 5.6 6.0 6.3 6.7 7.0
14. CaO outlet 125.9 135.6 145.2 154.7 164.3 173.9 183.5 193.1
15. CO2 outlet 123.7 132.9 142.7 152.1 161.3 170.6 180.1 189.9
Steam Slaker
16. CaO inlet 125.9 135.6 145.2 154.7 164.3 173.8 183.5 193.1
17. Steam, H2O inlet 54.1 54.1 54.1 54.1 54.1 54.1 54.1 54.1
18. Ca(OH)2 outlet 166.4 179.1 191.8 204.5 217.1 229.7 242.5 255.1

Table 5
Thermal and Electric Balance at different, climate-dependent CO2 capture rates.
Power [MW] CO2 Capture Rate

50 % 55 % 60 % 65 % 70 % 75 % 80 % 85 %

Thermal Power
Coolers − 89.3 − 94.7 − 98.6 − 103.7 − 107.9 − 111.3 − 116.4 − 120.6
Heaters 60.7 62.1 64.8 68.3 71.7 74.3 76.8 80.4
Steam Slaker − 37 − 42 − 47 − 52 − 62 − 62 − 68 − 72.6
Mix Tank − 62 − 62 − 63 − 63 − 63 − 63 − 63 − 64
Water Knockout − 9 − 11 − 12 − 14 − 17.5 − 18 − 19.5 − 21.6
Separator 2.56 2.53 2.5 2.45 2.42 2.38 2.35 2.31
Electrical Power
CO2 Compressor 14.9 15.85 17 18.3 19 20 21.6 23
Compressor – ASU 9.51 10.13 11 11.7 12.37 13.1 13.83 14.66
Compressor – NGCC 41.2 42.55 44 45.34 46.54 47.98 49.25 50.5
Gas Turbine − 75.01 − 77.23 − 79.9 − 82.36 − 84.31 − 89.46 − 89.46 − 92.15
Steam Turbine − 11.3 − 12 − 12.8 − 13.68 − 14.3 − 15.92 − 15.92 − 16.71
Auxiliary Equip. 20.7 20.7 20.7 20.7 20.7 20.7 20.7 20.7

Air Separation Unit (ASU), CO2 compression, and in the NGCC. The varies from 89.3 MW to 120.6 MW for the 85% CO2 capture rate. The
electrical power generation of the NGCC, including the steam and gas slaker is the fluidized bed reactor used to hydrate the quicklime (CaO)
turbine output, varies depending on the overall electrical power demand that comes from the calciner, and to recycle the Ca(OH)2 solution to the
of the system. The power from the NGCC depends on the mass flow rate pellet reactor. The steam slaker works at 300 ◦ C and atmospheric
entering the combined cycle. The details of the NGCC model and vali­ pressure, and the reaction specified in the block is the slaking reaction
dation details are described in the Supplementary file (Section 2). En­ between water and quicklime (CaO), where it enters at 652 ◦ C that
ergy requirements for the DAC plant are met by natural gas input to the generates the negative heat duty over varying CO2 capture rates.
calciner and NGCC. The energy consumption of the calciner changes
little with the CO2 capture rate as compared to the energy demand for 3.3. Water consumption
the NGCC which decreases from 5 to 3.1 GJ/tCO2 as the CO2 capture rate
increases due to the higher rate of increment on captured atmospheric The water demand in the direct air capture system depends on the
CO2 compared to the rate of natural gas input to combined cycle relative rates of water evaporation and generation in the air contactor,
(Figure S4). the scale of the type of process, and the mode of heat transfer for solvent
In the Calciner, 54% to 62% of total natural gas is fed to support the regeneration. Equations (1)–(3) are the representative reactions for CO2
calcining process to break down the CaCO3 to CaO and CO2. The highest absorption (along with a few other reactions listed in the supplementary
CO2 capture rate considered (85%) results in the largest CO2 flow, and file), and these show that water is also generated as a by-product. The
thus the largest energy requirement for each process. We assume a rate of water generation is based on the kinetics of absorption reactions,
constant auxiliary electric power demand of 20.7 MW (from Keith et al. which are dependent on the incoming air and solvent temperatures.
[10]), which accounts for the air contactor fans, pumping, and mixing With the variation of RH and temperature, the evaporation rate of water
power. The heating duty of the plant varies from 60.7 MW at a 50% also changes and that leads to the overall water makeup required. Fig. 3
capture rate to 80.36 MW at an 85% capture rate. Similarly, cooling duty represents the net water production/evaporation in the air contactors

6
K. An et al. Applied Energy 325 (2022) 119895

Fig. 3. (a) Net water production and (b) Net water production/consumption per tonne-CO2 in the Air-contactor at a wide range of climate conditions (negative sign
signifies evaporation; grey-colored area represents unreachable air conditions).

for various temperatures and RH. We estimate a net water consumption below 50 %, which may increase the energy demand (GJ/tCO2) signif­
of 6.6 t/tCO2 at 21 ◦ C and 69 % RH, which is somewhat higher than icantly. Therefore, while warmer temperatures are positive in acceler­
presented by Keith et al. (4.7 t/tCO2) [10]. Water loss from DAC has also ating the kinetics of CO2 capture, they may also increase water loss (at
been investigated by Rosa et al. [43], who estimated water loss from a lower relative humidity), which means that the trade-off between
high-temperature solvent-based DAC system between 0 t/tCO2 to 50 t/ climate (air temperature and relative humidity) and water availability
tCO2 as a function of air temperature, relative humidity, and molarity of will need to be carefully balanced.
the solvent.
Under hot and humid conditions, incoming air will lose sensible heat 3.4. Climate impact on CO2 capture efficiency and energy demand
based on the length of the air contactor over which air encounters sol­
vent. Water production is mainly due to Equation (2) and there is less The present model shows that the energy demand for industrial-scale
evaporation as it is already close to saturation. For instance, air can DAC could vary from 8.3 GJ/tCO2 to 11.1 GJ/tCO2 as a function of
reach a maximum RH of 100% at a saturation temperature of 32 ◦ C, and capture rate, which is determined by not only process design, but also
air will not hold any further moisture and any further increase of hu­ environmental conditions. The CO2 capture efficiency of the system is of
midity would result in precipitation or condensation. With an increase of importance, as it reflects the overall climate benefit of DAC. The CO2
temperature to 40 ◦ C, the maximum RH decreases to 63% (e.g., T:35 ◦ C, capture efficiency is considered as the ratio of the difference between the
maximum RH: 82%). The maximum RH achievable by the air for a CO2 captured and resulting GHG emissions (in CO2 equivalent terms) to
temperature of 40 ◦ C is 62%, which results in a net water loss of 736 t/h the captured CO2, and it is given as Equation (4).
or 5.42 t/tCO2 based on Fig. 3(b) [44]. In the present analysis, the inlet
temperature of air and solvent is assumed to be the same but, during the ṁCO2 ,captured − ṁCO2 ,NGCC − ṁCO2 ,grid − ṁCO2 ,upstream gas
ηCO2 ,capture = (4)
interaction of solvent and air along the length of the reactor, the heat of ṁCO2 ,captured
vaporization of moisture is taken from the sensible heat of liquid solvent,
where ṁCO2 ,captured is the capture rate of CO2 in tonnes from the at­
and that leads to decrease in solvent exit temperature.
For hot and dry conditions (e.g., RH of 20% at 40 ◦ C, which repre­ mosphere and the remaining terms represent emissions associated with
sents Arizona, US conditions in June), the humidity ratio in this condi­ DAC. Specifically, these are emissions from: integrated power genera­
tion is very low (9 g of moisture/kg of dry air), which leads to a very high tion, ṁCO2 ,NGCC , or any external sources of electricity, ṁCO2 ,grid ; and, natural
evaporation rate. The net water loss at this condition is 2340 t/h (19.8 t/ gas supply to the DAC facility (i.e. extraction, processing, and trans­
tCO2) at a CO2 capture rate of 83 %. Therefore, a large water make-up mission), ṁCO2 ,upstream gas . Strictly speaking, each of the emissions terms
needs to be arranged for DAC to operate under such climate condi­ should include climate impacts from their full life cycle (i.e., construc­
tions, and storing and pumping such large amounts of water would in­ tion, operation and decommissioning), but in the present analysis, only
crease the OPEX and auxiliary pumping power. It is worth noting that emissions associated with operation of the system are considered. This
the temperature has a larger influence on the water evaporation rate approach is consistent with that taken by previous researchers
than RH, as seen in Fig. 3. At 20% RH, as the inlet air temperature drops [17,29,45,46] who investigated the life cycle of DAC systems. Fig. 1
from 40 ◦ C to 0 ◦ C, the net water loss decreases from 2340 t/h (19.8 t/ shows that the electrical energy required to run the direct air capture
tCO2) to 650 t/h (7.3 t/tCO2) due to decrease saturation temperature plant is provided by an NGCC plant in the stand-alone case, the capture
and humidity ratio. For a higher RH condition, the humidity ratio results system for which is integrated with the DAC plant. For the stand-alone
in more evaporation, leading to increased KOH molarity and, thus, case, we include emissions from the NGCC plant (there is no imported
higher CO2 capture rates. This, in turn, results in formation of water as a electricity from the grid). To meet the second objective of the study, we
by-product compensating the water loss. For RH of 63% and if the substitute imported grid electricity for the NGCC plant and, in this case,
temperature drops from 40 ◦ C to 0 ◦ C, the net water loss decreases from we include emissions from external electricity generation (at an
736 t/h (5.42 t/tCO2) to 230 t/h (2.7 t/tCO2). From the above discus­ assumed grid intensity) since there is no electricity generation internal
sion, it can be concluded that net water loss is higher in hot and dry to the DAC plant.
conditions. It stands to reason that this will strongly influence deploy­ We consider a range of upstream emissions intensities for natural gas
ment of liquid solvent-based DAC in geographical locations with such supply to the DAC system. Three cases of natural gas upstream emissions
climate conditions. were considered from the most optimistic case, 3 gCO2e/MJ (at 0.3%
If DAC is deployed in a cold and dry location where the temperature CH4 emission rate), medium value, 15.5 gCO2e/MJ (at 1.57% CH4
drops well below 0 ◦ C, net water loss is lower as the effect of RH is emission rate) and a high value, 59.5 gCO2e/MJ (at 6% CH4 emission
smaller at lower temperatures. However, the CO2 capture rates drop rate) including pre-production, production, processing, transmission
using 100-year global warming potential (GWP 100) [47–49]. The

7
K. An et al. Applied Energy 325 (2022) 119895

resulting overall CO2 capture efficiency of the DAC system varies from climate conditions (T: 20 ◦ C RH: 60%). In contrast, CO2 capture effi­
32 % to 97 % as shown in Fig. 4. CO2 capture efficiency varies on two ciency would be around only 32% for the Alberta grid, which had a GHG
factors: CO2 capture rate and upstream emission of natural gas. In the intensity of 670 gCO2/kWh in 2019 [52]. The importance of choosing
the optimistic case, upstream emissions have a very minimal effect on carefully between energy supply options was highlighted by McQueen
CO2 capture efficiency of DAC plant over a wide range of CO2 capture et al. [19] given the impact of energy supply carbon intensity on CO2
rates. However, in the high case, the CO2 capture efficiency varies from capture efficiency [17,53]. Were the liquid-based DAC plant planned for
32 % to 49 % which is a wider range compared to both optimistic and the Permian Basin [54] in Texas to rely on grid electricity (with an
medium cases of natural gas upstream emissions. The energy con­ emissions intensity of around 335 gCO2/kWh [55]), the result would be a
sumption decreases exponentially from 11.1 to 8.3 GJ/tCO2 (a drop of CO2 capture efficiency of around 62%. Therefore, the idea of using grid
25%) as the CO2 capture rate increases from 40 % to 85%. This can be electricity instead of NGCC for solvent based DAC is highly debatable
compared to the energy consumption estimate from Carbon Engineering [19].
of 8.81 GJ/tCO2 at a capture rate of 75 % [10].
3.6. Climate impact on levelized cost of CO2 capture
3.5. Climate impact on CO2 capture efficiency in context
The climate impacts on the levelized cost of the DAC system are
We compare the sensitivity of CO2 removal efficiency (for both the analyzed for two scenarios: natural gas standalone and electricity grid-
natural gas standalone and grid-connected DAC system) to climate connected system. The equation for the levelized cost of CO2 capture
conditions with the carbon intensity of grid electricity, the global is provided as Equation (5).
warming potential time horizon, and the upstream methane emission CAPEX × CRF + O&MNon− Energy + O&MEnergy
rate. The results of the sensitivity analysis are shown in Fig. 5. In this Levelized Cost = U
(5)
Actual CO2 Capture Rate
work, we include CO2 emissions from the DAC plant flue gas (i.e.,
uncaptured CO2) and those resulting from natural gas supply because We use the baseline (75 % capture rate) levelized cost is from pre­
the climate only affects the operational emissions, not other emissions vious work [10], which includes capital cost, non-energy O&M, and
(e.g., in construction). The baseline CO2 capture efficiency of 86% is energy-related O&M. The capital cost and non-energy O&M (water,
based on 20 ◦ C air temperature, 60% relative humidity, and 1.57% labor, make-up chemicals) are normalized by the different amounts of
upstream methane emissions (15.5 g-CO2e/MJ using GWP 100) [49]. CO2 capture capacity that vary from different climate conditions.
Interestingly, the impact of climate on CO2 capture efficiency is small Energy-related O&M result from natural gas and electricity consumption
in both scenarios relative to the other factors shown in Fig. 5. While the derived from the Aspen process modeling. We use the capital cost re­
temperature and relative humidity variation result in a capture rate that ported by Keith et al. [10] for their natural gas standalone scenario (first-
varies between 55% and 92% at the air contactor, this has a relatively of-a-kind), updated from 2016 USD to 2021 USD (using the Chemical
small impact on the CO2 emission term in Equation (3). In contrast, Engineering Plant Cost Index [56]). Other economic parameters are
Fig. 5(a) shows that variation in methane emission rate between 6% and presented in Table 6, including a cost for water that could be sufficient to
0.2% results in a CO2 capture efficiency between 67% and 92%. The cover the losses due to evaporation from desalination in water-scarce
lower bound for the methane emission rate, 0.3% is a 2025 target set by locations ($0.67/m3 to $1.8/m3) [9]. We assume a capital recovery
the Oil and Gas Climate Initiative [50]. factor (CRF) of 12.5%, which is determined by the project cost of capital
For the electricity grid-connected system (i.e., no NGCC, but the and lifetime (e.g., 11.7% and 25 years).. The energy costs are assumed
calciner uses natural gas), the carbon capture efficiency is strongly $3.5/GJ for natural gas and $60/MWh for grid electricity [10].
dependent on the carbon intensity of electricity generation: variation in Annual average climate conditions (i.e. climate) are assumed to vary
electricity generation intensity from 10 to 680 g-CO2e/kWh results in from 10% to 90% of relative humidity for the temperature range from
variation in capture efficiency from 91% to 32% of CO2, as shown in 0 ◦ C to 40 ℃, excluding unreachable air conditions are considered.
Fig. 5(b) (see also Table S4). An electricity carbon intensity of 62 gCO2e/ Importantly, these climatic conditions reflect the average climate and
kWh is required to compete with the standalone NGCC case (for a cap­ temperature over the life of a DAC plant—daily and seasonal variations
ture efficiency of 86.3%), implying that there are few electricity grids in climate will lead to varying marginal costs of capture. For each cli­
from which it would make sense to draw electricity. For instance, with matic condition, the corresponding CO2 capture rates are calculated and
the 2020 Swedish grid intensity of 8.8 gCO2/kWh [51], a DAC plant used to estimate the levelized cost of CO2 capture from the atmosphere.
would have a carbon capture efficiency of close to 90 % under moderate The climate conditions affect not only the CO2 capture capacity from
different CO2 capture rates from the air contactor but also the natural
gas and electricity demand which were evaluated by the Aspen Plus
simulation. The levelized cost of CO2 capture, as shown in Fig. 6, in the
natural gas standalone scenario varies from $240/tCO2 to $409/tCO2
and the electricity grid-connected scenario varies from $265/tCO2 to
$440/tCO2. In both of these cases, the capital cost is consistently the
largest component of the calculated LCOE from 60.4% to 75.8%, fol­
lowed by non-energy O&M from 14.6% to 17.1%. For the energy-related
O&M, there are substantial difference for the grid connected system
(18.3% to 24.5%) and natural gas standalone (7.5% to 11.4%) mainly
due to the electricity cost (in this study, $60/MWh). For both scenarios,
the levelized cost is more sensitive to temperature than relative hu­
midity, which is consistent with earlier observations for the CO2 capture
rate (Fig. 2). It is, thus, clear that climate conditions significantly affects
the levelized cost for the DAC process, regardless of the energy supply.
This means that future assessments of DAC technology potential should
explicitly consider the regional climate when assessing the economic
potential of DAC.
Fig. 4. Energy consumption and CO2 capture efficiency over varying CO2
capture rates.

8
K. An et al. Applied Energy 325 (2022) 119895

Fig. 5. Sensitivity analysis on CO2 capture efficiency of (a) natural gas standalone and (b) electricity grid-connected system. Blue color is below the baseline and red
color is above the baseline).

tCO2 from 40% to 85% CO2 capture rate which is estimated by Aspen
Table 6 process modeling.
Economic parameters of capital and operational cost for DAC plant
The sensitivity analysis on the DAC system was performed for two
[10,56–59].
different systems, natural gas standalone and grid connected, and we
Project Year 25 varied the carbon intensity of grid electricity (for the grid connected
CRF (Capital Recovery Factor) [%/year] 12.5
system), global warming potential for methane, upstream methane
U (Utilization) [%] 90
Natural gas cost [$/GJ] 3.5 emission rate for natural gas, and different climate conditions. For both
Electricity cost [$/MWh] 60 scenarios, the climate impacts on CO2 capture efficiency are small
Water cost [$/m3] 1 relative to the impact of upstream natural gas supply GHG intensity for
CAPEX [$M] 1126.8
the stand-alone plant, and the carbon intensity of electricity generation
CEPCI Inflator (2016 to 2021) 1.35
Gross Product Inflator (2016 to 2021) 1.11 for the grid-connected plant. For example, varying the GHG intensity of
Labor Inflator (2016 to 2021) 1.16 electricity production from 10 to 680 gCO2e/kWh results in a reduction
of CO2 capture efficiency from 91% to 31.6%. Thus, from the standpoint
of climate benefits, fugitive emissions from the natural gas supply chain
4. Conclusions and carbon intensity of electricity are more important than the climate
in plant siting decisions.
The response of solvent-based DAC system was investigated at However, for both scenarios, we find that the levelized cost of CO2
different CO2 capture rates, which we show represent the performance capture can vary by a factor of nearly two between cold, dry conditions
of the system across a wide range of temperature and humidity condi­ and warm, humid conditions. We note that the levelized cost of CO2
tions. The air contactor was designed and simulated using Aspen soft­ capture is much more sensitive to temperature than relative humidity. In
ware to achieve the same CO2 capture rate, 75%, under reference addition, we estimate the levelized cost of CO2 capture for the natural
conditions presented by Keith et al. [10], and the climate conditions gas standalone DAC system to be 7% to 10% lower than the electricity
(specifically temperature and RH) were varied to estimate the actual grid-connected scenario for all ranges of climate conditions. The lev­
CO2 capture rate. The CO2 capture efficiency was estimated for both the elized cost of CO2 direct air capture for a liquid solvent-based system
natural gas standalone and grid-electricity connected DAC systems. The can, thus, be reduced by deploying the plant in a warm and humid
energy demand of the DAC plant was found to vary from 8.3 to 11.1 GJ/ climate which maximizes the CO2 capture rate at the air contactor

9
K. An et al. Applied Energy 325 (2022) 119895

Fig. 6. Levelized cost of DAC powered by (a) natural gas standalone and (b) electricity grid-connected system (Grey-colored area represents unreachable
air conditions).

(maximizing capital utilization) and reduces the demand for makeup Appendix A. Supplementary material
water and requires a lower energy consumption (minimizing operating
costs). Supplementary data to this article can be found online at https://doi.
The objective of the study was to assess the impact of climate con­ org/10.1016/j.apenergy.2022.119895.
ditions (temperature and RH) on liquid-solvent based DAC systems and
how it might influence the CO2 capture rates, CO2 capture efficiency, References
and cost of DAC. The results show that deploying DAC in some locations
will result in a lower cost than in others simply by virtue of the climate [1] Eyring V, Gillett NP, Achuta Rao KM, Barimalala R, Barreiro Parrillo M, Bellouin N,
et al. Human Influence on the Climate System. In: Valérie M-D, Zhai P, Pirani A,
conditions. The implications of this study are multiple. First, any eco­ Connors S, Pean C, Berger S, et al., editors. Climate Change 2021: The Physical
nomic assessment of DAC potential should consider specific climate Science Basis. Contribution of Working Group I to the Sixth Assessment Report of
conditions rather than just assuming generic performance figures at the Intergovernmental Panel on Climate Change, Cambridge University Press; In
Press.
moderate temperature and relative humidity conditions. Second, the [2] Seneviratne SI, Zhang X, Adnan M, Badi W, Dereczynski C, Di Luca A, et al.
results suggest that climate conditions should be added to the list of Weather and Climate Extreme Events in a Changing Climate. In: Valérie M-D, Zhai
factors—such as proximity to geologic CO2 storage and access to ener­ P, Pirani A, Connors S, Pean C, Berger S, et al., editors. Climate Change 2021: The
Physical Science Basis. Contribution of Working Group I to the Sixth Assessment
gy—that influence the deployment of DAC facilities and, indeed, Report of the Intergovernmental Panel on Climate Change, Cambridge University
emphasize emerging findings that geographical location matters when Press; In Press.
siting DAC facilities cannot be located anywhere. Finally, it is likely that [3] Hilaire J, Minx JC, Callaghan MW, Edmonds J, Luderer G, Nemet GF, et al.
Negative emissions and international climate goals—learning from and about
the solid-sorbent based DAC systems will also exhibit a sensitivity to
mitigation scenarios. Clim Change 2019;157(2):189–219.
climate conditions, which means that different types of DAC systems [4] Smith P, Davis SJ, Creutzig F, Fuss S, Minx J, Gabrielle B, et al. Biophysical and
could be optimal in specific locations. economic limits to negative CO2 emissions. Nature Clim Change 2016;6(1):42–50.
[5] Minx JC, Lamb WF, Callaghan MW, Fuss S, Hilaire J, Creutzig F, et al. Negative
emissions—Part 1: Research landscape and synthesis. Environ Res Lett 2018;13(6):
CRediT authorship contribution statement 063001. https://doi.org/10.1088/1748-9326/aabf9b.
[6] Realmonte G, Drouet L, Gambhir A, Glynn J, Hawkes A, Köberle AC, et al. An inter-
Keju An: Conceptualization, Methodology, Investigation, Visuali­ model assessment of the role of direct air capture in deep mitigation pathways. Nat
Commun 2019;10:3277. https://doi.org/10.1038/s41467-019-10842-5.
zation, Writing – original draft. Azharuddin Farooqui: Methodology, [7] Nemet GF, Callaghan MW, Creutzig F, Fuss S, Hartmann J, Hilaire J, et al. Negative
Investigation, Writing – review & editing. Sean T. McCoy: Conceptu­ emissions—Part 3: Innovation and upscaling. Environ Res Lett 2018;13(6):063003.
alization, Supervision, Writing – review & editing, Funding acquisition. https://doi.org/10.1088/1748-9326/aabff4.
[8] Breyer C, Fasihi M, Aghahosseini A. Carbon dioxide direct air capture for effective
climate change mitigation based on renewable electricity: a new type of energy
Declaration of Competing Interest system sector coupling. Mitig Adapt Strateg Glob Change 2020;25:43–65. https://
doi.org/10.1007/s11027-019-9847-y.
[9] Fasihi M, Efimova O, Breyer C. Techno-economic assessment of CO2 direct air
The authors declare that they have no known competing financial capture plants. J Cleaner Prod 2019;224:957–80. https://doi.org/10.1016/j.
interests or personal relationships that could have appeared to influence jclepro.2019.03.086.
the work reported in this paper. [10] Keith DW, Holmes G, St. Angelo D, Heidel K. A Process for Capturing CO2 from the
Atmosphere. Joule 2018;2(8):1573–94.
[11] Beuttler C, Charles L, Wurzbacher J. The Role of Direct Air Capture in Mitigation of
Data availability Anthropogenic Greenhouse Gas Emissions. Front Clim 2019;1:10. https://doi.org/
10.3389/fclim.2019.00010.
Data will be made available on request. [12] E. Bajamundi CJ, Koponen J, Ruuskanen V, Elfving J, Kosonen A, Kauppinen J,
et al. Capturing CO2 from air: Technical performance and process control
improvement. J CO2 Util 2019;30:232–9.
Acknowledgements [13] Gutknecht V, Snæbjörnsdóttir SÓ, Sigfússon B, Aradóttir ES, Charles L. Creating a
carbon dioxide removal solution by combining rapid mineralization of CO2 with
direct air capture. Energy Procedia 2018;146:129–34. https://doi.org/10.1016/j.
This research was undertaken thanks in part to funding from the egypro.2018.07.017.
Canada First Research Excellence Fund. [14] Didas SA, Choi S, Chaikittisilp W, Jones CW. Amine-Oxide Hybrid Materials for CO
2 Capture from Ambient Air. Acc Chem Res 2015;48:2680–7. https://doi.org/
10.1021/acs.accounts.5b00284.

10
K. An et al. Applied Energy 325 (2022) 119895

[15] Stampi-Bombelli V, van der Spek M, Mazzotti M. Analysis of direct capture of $ [36] Pina-Martinez A, Privat R, Jaubert J-N, Peng D-Y. Updated versions of the
${\hbox {CO}}_{2}$$ from ambient air via steam-assisted temperature–vacuum generalized Soave α-function suitable for the Redlich-Kwong and Peng-Robinson
swing adsorption. Adsorption 2020;26:1183–97. https://doi.org/10.1007/s10450- equations of state. Fluid Phase Equilib 2019;485:264–9. https://doi.org/10.1016/
020-00249-w. j.fluid.2018.12.007.
[16] Choi S, Drese JH, Eisenberger PM, Jones CW. Application of Amine-Tethered Solid [37] Sanku MG, Svensson H. Modelling the precipitating non-aqueous CO2 capture
Sorbents for Direct CO 2 Capture from the Ambient Air. Environ Sci Technol 2011; system AMP-NMP, using the unsymmetric electrolyte NRTL. Int J Greenhouse Gas
45:2420–7. https://doi.org/10.1021/es102797w. Control 2019;89:20–32. https://doi.org/10.1016/j.ijggc.2019.07.006.
[17] Deutz S, Bardow A. Life-cycle assessment of an industrial direct air capture process [38] Kim J, Na J, Kim K, Bak JH, Lee H, Lee U. Learning the properties of a water-lean
based on temperature–vacuum swing adsorption. Nat Energy 2021;6:203–13. amine solvent from carbon capture pilot experiments. Appl Energy 2021;283:
https://doi.org/10.1038/s41560-020-00771-9. 116213. https://doi.org/10.1016/j.apenergy.2020.116213.
[18] Holmes G, Nold K, Walsh T, Heidel K, Henderson MA, Ritchie J, et al. Outdoor [39] Rastegar Z, Ghaemi A. CO2 absorption into potassium hydroxide aqueous solution:
Prototype Results for Direct Atmospheric Capture of Carbon Dioxide. Energy experimental and modeling. Heat Mass Transf 2022;58:365–81. https://doi.org/
Procedia 2013;37:6079–95. 10.1007/s00231-021-03115-9.
[19] McQueen N, Desmond MJ, Socolow RH, Psarras P, Wilcox J. Natural Gas vs. [40] Sander R. Compilation of Henry’s law constants (version 4.0) for water as solvent.
Electricity for Solvent-Based Direct Air Capture. Front Clim 2021;2:618644. Atmos. Chem Phys 2015;15:4399–981. https://doi.org/10.5194/acp-15-4399-
https://doi.org/10.3389/fclim.2020.618644. 2015.
[20] Holmes G, Keith DW. An air–liquid contactor for large-scale capture of CO 2 from [41] Cadogan SP, Maitland GC, Trusler JPM. Diffusion Coefficients of CO 2 and N 2 in
air. Phil Trans R Soc A 2012;370:4380–403. https://doi.org/10.1098/ Water at Temperatures between 298.15 K and 423.15 K at Pressures up to 45 MPa.
rsta.2012.0137. J Chem Eng Data 2014;59:519–25. https://doi.org/10.1021/je401008s.
[21] Wilcox J, Rochana P, Kirchofer A, Glatz G, He J. Revisiting film theory to consider [42] Haghighi H, Chapoy A, Tohidi B. Freezing Point Depression of Electrolyte
approaches for enhanced solvent-process design for carbon capture. Energy Solutions: Experimental Measurements and Modeling Using the Cubic-Plus-
Environ Sci 2014;7:1769. https://doi.org/10.1039/c4ee00001c. Association Equation of State. Ind Eng Chem Res 2008;47:3983–9. https://doi.org/
[22] McQueen N, Psarras P, Pilorgé H, Liguori S, He J, Yuan M, et al. Cost Analysis of 10.1021/ie800017e.
Direct Air Capture and Sequestration Coupled to Low-Carbon Thermal Energy in [43] Rosa L, Sanchez DL, Realmonte G, Baldocchi D, D’Odorico P. The water footprint of
the United States. Environ Sci Technol 2020;54(12):7542–51. carbon capture and storage technologies. Renew Sustain Energy Rev 2021;138:
[23] Sabatino F, Grimm A, Gallucci F, van Sint AM, Kramer GJ, Gazzani M. 110511. https://doi.org/10.1016/j.rser.2020.110511.
A comparative energy and costs assessment and optimization for direct air capture [44] Ayoub A, Gjorgiev B, Sansavini G. Cooling towers performance in a changing
technologies. Joule 2021;5:2047–76. https://doi.org/10.1016/j. climate: Techno-economic modeling and design optimization. Energy 2018;160:
joule.2021.05.023. 1133–43. https://doi.org/10.1016/j.energy.2018.07.080.
[24] Keith DW, Ha-Duong M, Stolaroff JK. Climate strategy with CO2 capture from the [45] de Jonge MMJ, Daemen J, Loriaux JM, Steinmann ZJN, Huijbregts MAJ. Life cycle
air. Clim Change 2006;74:17–45. https://doi.org/10.1007/s10584-005-9026-x. carbon efficiency of Direct Air Capture systems with strong hydroxide sorbents. Int
[25] Sanz-Pérez ES, Murdock CR, Didas SA, Jones CW. Direct Capture of CO 2 from J Greenhouse Gas Control 2019;80:25–31. https://doi.org/10.1016/j.
Ambient Air. Chem Rev 2016;116:11840–76. https://doi.org/10.1021/acs. ijggc.2018.11.011.
chemrev.6b00173. [46] Deutz S, Bardow A. How (carbon) negative is direct air capture? Life cycle
[26] Larsen J, Herndon W, Grant M, Marsters P. Capturing Leadership: Policies for the assessment of an industrial temperature- vacuum swing adsorption process.
US to Advance Direct Air Capture Technology. Washington, D.C.: Rhodium Group, ChemRxiv 2020:28.
LLC; 2019. [47] Liu RE, Ravikumar AP, Bi XT, Zhang S, Nie Y, Brandt A, et al. Greenhouse Gas
[27] Lackner KS, Azarabadi H. Buying down the Cost of Direct Air Capture. Ind Eng Emissions of Western Canadian Natural Gas: Proposed Emissions Tracking for Life
Chem Res 2021;60:8196–208. https://doi.org/10.1021/acs.iecr.0c04839. Cycle Modeling. Environ Sci Technol 2021;55(14):9711–20.
[28] Meckling J, Biber E. A policy roadmap for negative emissions using direct air [48] Bauer C, Treyer K, Antonini C, Bergerson J, Gazzani M, Gencer E, et al. On the
capture. Nat Commun 2021;12:2051. https://doi.org/10.1038/s41467-021- climate impacts of blue hydrogen production. Sustain Energy Fuels 2021;6(1):
22347-1. 66–75.
[29] Madhu K, Pauliuk S, Dhathri S, Creutzig F. Understanding environmental trade-offs [49] Littlefield JA, Marriott J, Schivley GA, Skone TJ. Synthesis of recent ground-level
and resource demand of direct air capture technologies through comparative life- methane emission measurements from the U.S. natural gas supply chain. J Cleaner
cycle assessment. Nat Energy 2021;6:1035–44. https://doi.org/10.1038/s41560- Prod 2017;148:118–26. https://doi.org/10.1016/j.jclepro.2017.01.101.
021-00922-6. [50] Cooper J, Balcombe P, Hawkes A. The quantification of methane emissions and
[30] Terlouw T, Bauer C, Rosa L, Mazzotti M. Life cycle assessment of carbon dioxide assessment of emissions data for the largest natural gas supply chains. J Cleaner
removal technologies: a critical review. Energy Environ Sci 2021;14:1701–21. Prod 2021;320:128856. https://doi.org/10.1016/j.jclepro.2021.128856.
https://doi.org/10.1039/D0EE03757E. [51] European Environment Agency. Greenhouse gas emission intensity of electricity
[31] Liu CM, Sandhu NK, McCoy ST, Bergerson JA. A life cycle assessment of generation in Europe 2021.
greenhouse gas emissions from direct air capture and Fischer-Tropsch fuel [52] NIR. NATIONAL INVENTORY REPORT 1990 –2019: GREENHOUSE GAS SOURCES
production. Sustain Energy Fuels 2020;4:3129–42. https://doi.org/10.1039/ AND SINKS IN CANADA 2019.
C9SE00479C. [53] Viebahn P, Scholz A, Zelt O. The Potential Role of Direct Air Capture in the German
[32] Pinsent BRW, Pearson L, Roughton FJW. The kinetics of combination of carbon Energy Research Program—Results of a Multi-Dimensional Analysis. Energies
dioxide with hydroxide ions. Trans Faraday Soc 1956;52:1512. https://doi.org/ 2019;12:3443. https://doi.org/10.3390/en12183443.
10.1039/tf9565201512. [54] Carbon Engineering. Engineering of world’s largest Direct Air Capture plant begins
[33] Kucka L, Kenig EY, Górak A. Kinetics of the Gas− Liquid Reaction between Carbon 2019.
Dioxide and Hydroxide Ions. Ind Eng Chem Res 2002;41:5952–7. https://doi.org/ [55] Austin Energy. Austin Energy system average carbon intensity. 2015. doi:
10.1021/ie020452f. 10.26000/040.000019.
[34] Garcia M, Knuutila HK, Gu S. ASPEN PLUS simulation model for CO 2 removal with [56] Vatavuk WM. Updating the CE plant cost index. Chem Eng 2002;109:62–70.
MEA: Validation of desorption model with experimental data. J Environ Chem Eng [57] Ray Sinnott, Gavin Towler. Chemical engineering design: principles, practice and
2017;5:4693–701. https://doi.org/10.1016/j.jece.2017.08.024. economics of plant and process design. 2021.
[35] Liheng G, Yudong D, Xiaoqiang Li, Xun Zu, Qiang L, Shaojun Y. Simulation and [58] Chemical Engineering Plant Cost Index, Chemical Engineering essentials for the
optimization study on aqueous mea-based co2 capture process. Chem Eng Trans CPI professional 2021.
2018;70:751–6. https://doi.org/10.3303/CET1870126. [59] Palligarnai T. Vasudevan, Gael D. Ulrich. How to Estimate Utility Costs.
CHEMICAL ENGINEERING 2006:4.

11

You might also like