Specification of The ECT Framework Methodologies For Direct Air Capture

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 43

Specification

of the ECT Framework


methodologies
for
Direct Air Capture

V1.0
25th August 2021

1
1
Contents

Contents ............................................................................................................................... 2
About this document ............................................................................................................. 3
Technology description ......................................................................................................... 4
Research methodology ...................................................................................................... 4
Technology pathways ........................................................................................................ 5
High Temperature Direct Air Capture (HT-DAC) ............................................................ 6
Low Temperature Direct Air Capture (LT-DAC) .............................................................. 7
Market uptake model............................................................................................................. 9
Role of policies and public support in the normal and accelerated deployment scenarios
.................................................................................................................................... 12
Incumbent and alternative technologies ........................................................................... 14
System boundary ............................................................................................................ 14
LCA Approach ............................................................................................................. 16
Reference scenario ......................................................................................................... 16
GHG emission rates ........................................................................................................ 16
Incumbent .................................................................................................................... 16
Alternative technologies ............................................................................................... 17
Direct and indirect emissions: carbon storage and utilization ....................................... 18
GHG impact assessment method ................................................................................ 20
Carbon Abatement factor................................................................................................. 21
Emission Reduction Potential .......................................................................................... 22
Avoided emissions........................................................................................................... 23
Uncertainty and challenges ............................................................................................. 23
Uncertainty related to methodology .............................................................................. 23
Uncertainty related to GHG data .................................................................................. 24
Challenges related to secondary enabling and rebound effects ................................... 25
Reductions in Green Premium ............................................................................................ 27
Learning curves ............................................................................................................... 28
Cost models ................................................................................................................. 29
Uncertainty and challenges ............................................................................................. 35
Catalyzed emissions reductions .......................................................................................... 36
Uncertainty and challenges ............................................................................................. 38
References ......................................................................................................................... 39

2
2
About this document
This document contains a technology-specific methodology for Direct Air Capture (DAC) in
line with the Emerging Climate Technology (ECT) Framework version 1.0. This methodology
specifies the market uptake model and calculation of ex-ante metrics – Emission Reduction
Potential, reductions in Green Premium and Catalyzed Emissions Reductions (CatER) - for
the following DAC production technology pathways:
- High Temperature DAC (HT-DAC), causticization with alkali hydroxides
- Low Temperature DAC (LT-DAC), Organic-inorganic hybrid sorbents (amines in
porous adsorbents)

The key input parameters such as carbon abatement factors, market size, and market
dynamics are based on global average values and models and are provided specifically for
ex-ante estimation of systemic effects of the acceleration of the technology deployment
globally. For ex-post estimate of avoided emissions from individual projects, project-specific
data and local emission factors should be applied, which may differ from the average values
provided here.
The use of scenarios for forward-looking projections come with inherent uncertainties driven
by imperfect knowledge and assumption biases (related to climate, market, socio-economic
and technological developments), and uncertainties in statistics where limited data is
available. Scenario sensitivity analysis can help in identifying the most representative
pathways and parameters, applicable to specific use-cases. Minimum and maximum values
for some input values are provided to demonstrate the possible range and the inherent
uncertainty of the calculated figures.
The research work has been conducted CDP and Breakthrough Energy during the first half of
2021.

3
3
Technology description
This chapter presents the methodology used to collect data and information about DAC
technology followed by the description of each relevant technology pathway.

Research methodology
The research was conducted by a desk analysis of the scientific and technical literature related
to the Direct Air Capture technology and the low-temperature and high-temperature DAC
pathways. The scope of the review has included life cycle-based studies, technical reports
(to capture climate related life cycle impacts) and capital, operational and product cost and
market assessments (to gauge Green Premium and market diffusion rates) from academic
and industry researchers, technology developers, and international agencies. Furthermore,
semi-structured interviews were conducted with academics, International Energy Agency
(IEA) representatives and technology developers.
Overall, 62 papers on DAC technology were reviewed. The papers used in the analysis were
selected based on the following criteria:

 Transparency: Providing clear information about the key methodological assumptions


underlying the study (e.g., technology pathways, data quality, completeness of the life
cycle inventory, and system boundaries)
 Provision of up-to-date information: Papers published between 2015 and 2021
were preferred to ensure the findings represent the current state of the art. A few older
papers were used to ensure all relevant information and technology improvements are
captured.
As for the greenhouse gas (GHG) assessment the following information has been assessed:

 GHG emission rates of each identified technology and pathway.


 Analysis of the major life cycle stages and emissions contributors.
 Direct and indirect (upstream and downstream) emissions.
 Uncertainties and critical parameters of the calculation.
 Comparison with the BAU scenario with the associated GHG emissions (for avoided
emissions calculations).
In general, for the DAC technologies, cradle-to-grave LCA studies has been used, assessing
all the production stages and the end-of-life phase emissions of the technologies. The
variability of the results come from the differences in technology configurations, project
parameters, differences in the upstream and direct impacts (e.g., energy source for the
technologies).
In terms of economic assessment (for Green Premium and diffusion calculations), the
following has been evaluated:

 Market size currently and potential market size changes.


 Market diffusion rates of the technologies.
 Current technology costs (capital costs, operational costs, product and production
costs).
 Technology readiness levels (to gauge the learning potential and risk of technology
failure).
 Learning rates of the technologies and potential technology cost impacts (forward
looking).
4
4
Regarding top-level economic indicators and market trends, established modelling agency
outputs (e.g., IEA, IPCC) and models were used as guidelines. For technology and project
specifics, inputs from technology developers were considered alongside academic
assessments.

Technology pathways
Direct air capture (DAC) is one of several carbon removal options available for anthropogenic
GHG emissions reduction. Unlike nature-based solutions (e.g., reforestation and coastal
habitat restoration) or natural process enhancement (e.g., increasing the carbon content of
the soil through land management), DAC is a technology-based solution, and is similar to point
source capture technologies such as bioenergy with carbon capture and storage (BECCS).
DAC is an undoubtedly promising approach to atmospheric Carbon Dioxide Removal (CDR),
although it can only become climate relevant if its deployment reaches the gigaton scale by
the middle of this century (Beuttler et al., 2019). One clear advantage of the technology is that
it can help manage emissions from hard-to-abate sectors, such as heavy industries and
aviation, and on account of that, DAC can support a faster transition1.
DAC is not an alternative to cutting emissions or an excuse for delayed action, as relying on
this technology alone can lead to a global temperature overshoot of up to 0.8 °C. Nonetheless,
the technology can play an important role in achieving climate goals (Realmonte et al., 2019).
To realize the ambitious 20% annual capacity growth considered in some mitigation scenarios,
large-scale demonstrations of DAC technologies will be needed. This deployment rate (I.e.,
about 1,500 megaton scale facility each year between 2030-2050) is challenging but it is not
outside the ranges of other technological revolutions, such as wind turbines, solar
photovoltaics or jet engines. Strong targeted government2 and policy support, and substantial
investment3 through grants, tax credits (e.g., 45Q in the US), as well as public procurement of
CO2 offsets, in turn could boost further research and development, facilitate deployment, and
learning through operations (Gambhir & Tavoni, 2019). 
Additionally, any large-scale deployment of the technology requires further research to fully
understand its system-level impacts, such as ocean outgassing, intergenerational implications
(potential delayed action, intergenerational equity, moral hazard), impacts on the chemicals
industry (NaOH, monoethylamine (MEA) sorbent production), and the energy consumption of
the full DAC value chain, from sorbent production to operation of the fans (Gambhir & Tavoni,
2019), (Chatterjee et al., 2020).
DAC is a technology to capture CO2 directly from the ambient air and to generate a
concentrated stream of CO2. The captured CO2 can either be stored in deep geological
formations (permanent removal) or used in the production of fuels, chemicals, building
materials and other products containing CO 2. The two different pathways for the captured CO 2
require different CO2 concentrations, e.g., sequestration in a geological storage requires pure
CO2 products (concentration > 99%), whereas, for instance, utilization in agriculture can
function with more dilute products (~ 5%). This can have substantial impact on the cost of
production and the price of the final CO2 product. As CO2 needs to be compressed at a very

1
https://www.iea.org/reports/direct-air-capture
2
E.g., the U.S. Department of Energy (DOE) on the 5th of March 2021 announced up to $24 million for
research into the DAC technology.
3
On the 2nd of June 2021, the European Commission and Breakthrough Energy Catalyst announced
a new partnership with the aim to mobilise new investments of up to €820 million/$1 billion between
2022-26 to build large-scale, commercial demonstration projects for clean technologies.
5
5
high pressure in order to be injected into geological formations, the compression step naturally
increases the energy need, capital cost (due to the requirement for additional equipment such
as a compressor), the operating costs of the plant (such as running the compressor) and as a
result the cost of capture. That said, the quality and quantity of energy needed, and costs vary
not only according to the final utilization of the captured CO2 but also according to the type of
technology used, i.e., solid or liquid sorbents.
All the commercial DAC technologies roughly follow the below steps:

 Large fans push ambient air through a filter.


 The ambient air comes into contact with chemical media (e.g., aqueous alkaline
solvent or functionalized sorbent).
 The chemical media are subsequently stripped of CO2 through the application of (heat)
energy.
 This results in a CO2 stream that typically undergoes dehydration and compression.
 The chemical media is regenerated for reuse.
Although multiple technology pathways are available to capture CO2 from the air, the most
common and mature processes are the liquid and solid direct air capture systems. In the
former, air passes through chemical solutions (e.g., a hydroxide solution) that remove the CO2;
in the latter, a solid sorbent which chemically binds with CO2 is heated up which causes the
release of concentrated CO2 that is captured.

High Temperature Direct Air Capture (HT-DAC)

The High Temperature DAC (HT-DAC) refers to a variety of designs where the ambient air is
pushed through a liquid solution and the CO2 is captured through absorption or adsorption
with a sorbent or solvent. The main drawback of the process is the high thermal energy need,
as the captured CO2 must be separated from the precipitate. The different methods employ
different chemicals to manage the energy need.
One of the most promising of the HT-DAC systems uses aqueous hydroxide sorbents
(solutions of NaOH, KOH). This design was proposed by the American Physical Society (APS)
in 2011 (Socolow et al., 2011) and further developed through the work of Keith and Holmes
(Keith et al., 2018) in collaboration with Climate Engineering.

Similarly to other HT-DAC systems, the hydroxide solutions require high-temperature heat to
be regenerated (T > 800 °C). Although the technology is more mature than solid sorbent
capture and the design allows for large scale plants, capturing 1 MtCO 2/year requires major
capital investments (for building plant facilities) with limited potential for future cost reduction.
Moreover, with the current design 5 to 13 metric tons of water is used for each metric ton of
carbon dioxide captured during normal operation, depending on humidity and air temperature
(Realmonte et al., 2019). According to Keith (2018), under ideal operating atmospheric
conditions (20°C and 64% relative humidity), the plant needs 4.7 tons of water per ton CO2
captured. A two-loop hydroxide-carbonate system (NaOH-CaOH or KOH-CaOH), with a plant
reference size of 1 Mt/year of captured CO2 requires between 1.3 to 1.8 GJ/tCO2 electricity
and heat input between 5.3 to 8.1 GJ/tCO2 (Socolow et al., 2011 and Keith et al., 2018). See
the systems schematic in Figure 1., below.

6
6
Figure 1 - High temperature DAC technology process flowchart

Some of the considerable system design benefits are because it can be built using cheap
cooling-tower hardware, the (liquid) surface is continuously renewed allowing very long
contactor lifetimes despite dust and atmospheric contaminants, CO2 can be easily pumped to
a central regeneration facility allowing economies of scale. However, the cost and complexity
of the regeneration system and water loss in dry environments are considered disadvantages
(Keith et al., 2018). Furthermore, this design also demands large amounts of energy, mainly
high-temperature heat for sorbent regeneration. As the heat energy can be provided by on-
site natural gas combustion an additional 95% capture efficiency carbon capture and storage
(CCS) unit is needed to capture the CO2 emitted, this increases the complexity and capital
costs of the design.

Low Temperature Direct Air Capture (LT-DAC)

The Low Temperature DAC (LT-DAC) refers to a highly versatile procedure in direct air
capture where a chemical species that chemically bind with CO2 is deposited into the pores or
on the surface of a solid support. After the CO2 is adsorbed from ambient air, the filters are
heated releasing the concentrated CO2 which can be captured for storage or use.
The most studied chemicals for the technology are amine adsorbents (e.g.,
monoethanolamine (MEA)) that require only approximately 85–120 °C for desorption offering
the possibility of low energy input (waste heat can be used in the process). The solid sorbent
capture is a less mature modular design, that might be suitable for mass production and
potentially rapid cost reduction (low level of technological maturity). Although capital costs
might be smaller than the HT-DAC design, sorbent degradation may lead to high operational
expenditure. Hence, for solid sorbent chemisorption low volatility amine species have been
studied to limit sorbent degradation and amine loss (LT-DAC amine efficiency of 0.10 to 0.14
mol of CO2/mol of N). Furthermore, the system is less susceptible than HT-DAC to changes
in water humidity, there is little difference between dry and wet air performance.
Due to commercial confidentiality the information on the system is limited. Solid sorbents have
lower energy consumption, given the lower regeneration temperature: electricity use is
between 0.6 to 1.1 GJ/tCO2; and heat between 4.4 to 7.2 GJ/tCO2 (Ishimoto et al., 2017).
Water consumption does not seem to be a concern; a typical plant using solid sorbent
adsorption is estimated to use 1.6 metric tons of water for each metric ton of CO2 captured
(NASEM, 2019), Realmonte et al., 2019). See the systems schematic in Figure 2 below.

7
7
Figure 2- Low temperature DAC technology process flowchart

The below table (Table 1) provides a summary of the energy use of direct air capture. Note,
the price floor of the two technologies are estimated using optimized plant designs (e.g., by
Keith for HT-DAC) and in the case of LT-DAC by the moisture-swing technology which
drastically reduces energy needs and final costs as it uses natural airflows (wind-speed) to
move air around the filters.

Technology Electricity [GJ/tCO2] Heat [GJ/tCO2]


HT-DAC High 1.8 8.1
Low 1.3 5.3
LT-DAC High 1.1 7.2
Low 0.6 4.4
Table 1 - DAC technology pathways energy requirements (Realmonte et al., 2019)

8
8
Market uptake model
The total global available capacity by 2050 for carbon capture was identified as 7,678 million
metric tons of CO2 (Fasihi et al., 2019). The alternative to DAC - and Negative Emission
Technologies (NETs) in general - is more stringent mitigation now. In other words, the analysis
of future scenarios tells us that future market share of DAC is closely associated with the need
for its existence due to overshooting, expectations on mitigation costs and costs of alternative
technologies (e.g. BECCS). Figure 3 shows the total capacity that could be technically
available over time as per Fasihi (2019)4, contrasted with the scenarios used.
The Net Zero model is ambitious from the point of view of mitigation, but still points to
approximately 0.7 Gt of CO2 capture with DAC, far from the available capacity provided by
Realmonte, (2019) and Fasihi (2019). In the IEA model it is assumed that around 95% of total
CO2 captured in 2050 is stored in permanent geological storage and 5% is used to provide
synthetic fuels, therefore in all our estimations (including market penetration, GHG intensities,
etc.) only direct air carbon capture with storage (DACCS) was used. For the reference market
uptake model, a middle-way scenario for DAC has been constructed using an ambitious
scenario to meet 1.5ºC (IEA Net Zero by 20505). To achieve a middle-way scenario for DAC
deployment, the 2050 Net Zero Emissions scenario (NZE) capacity and carbon price is shifted
to 2095, with interim deployment estimates also shifted based on interim carbon prices. The
reference scenario constructed herein does not have a target net zero year but uses social
cost of carbon as the carbon price pathway6. This method yields only a few (i.e., four) market
penetration point estimates Table 2 and Figure 4.

DAC
Year (MtCO2)
2020 0.0055
2046 72
2080 528
2095 633
Table 2 - Data used as reference scenario for total capacity for DAC, in accordance with IEA ETP RTS
2020 and IEA Net Zero by 2050 report

4
With the following modification: for the 2020 starting point the actual DAC capacity value provided by
the IEA was considered and the 2020 demand baseline by Fasihi was pushed back to 2025 in line with
the IEA’s records on existing DAC capacity.
5
https://www.iea.org/reports/net-zero-by-2050
6
https://www.epa.gov/sites/default/files/2016-12/documents/social_cost_of_carbon_fact_sheet.pdf
9
9
Figure 3 –Total technical capacity for DAC until 2050 (Fasihi, 2019) relative to DAC scenarios

DAC scenarios
1000
DAC installed capacity (MMTCO2)

900
800
700
600
500
400
300
200
100
0
2020 2030 2040 2050 2060 2070 2080 2090 2100

Reference Scenario NZE by 2050

Figure 4– Detailed view of DAC technology uptake scenarios. The Reference scenario is defined by the
market penetration point estimates between 2020-20957,8

To be able to estimate the impacts of investments on the baseline for every year until 2050, a
curve was fitted to the market penetration points. For the curve fit, we use the Bass function,
where the key parameters of imitator and innovator constant are calculated by fitting the curve
to the values in 2046, 2080 and 2095 using ordinary least squares (OLS) regression. The
Bass equation as follows:

Where:

7
The NZE by 2050 scenario is as specified by the IEA, https://www.iea.org/reports/net-zero-by-2050
8
In the Reference scenario plot 95% of the total DAC market is considered, as per the NZE scenario,
that assumes around 95% of total CO2 captured is stored permanently and 5% is used to produce
synthetic fuels. Our assessment considered DACCS only.
10
10
λ: is the learning coefficient
η: is the demand elasticity
m: is the total market size targeted, anticipated, or possible
α: is the innovator coefficient, describing market dynamics near T=0
β: is the imitator coefficient, describing later market dynamics
Q: is the total experience (as quantity)
This market uptake scenario is then accelerated by catalytic investments in the catalyzed
scenario (Figure 5.

Figure 5 – Bass diffusion curve for baseline market penetration and for the accelerated scenario between
2020 and 2050 for the HT-DAC pathway

Market share was then allocated equally between the two DAC pathways. Although the HT-
DAC pathway is more mature and large-scale removal facilities are closer to completion, it is
the more capital and energy intensive pathway. LT-DAC is the less mature of the two
technologies, therefore it is likely to go through a steeper learning and cost-reduction curve,
additionally the LT-DAC pathway requires less energy and is considered more modular. The
pathway specific trajectory is shown in Table 3. Please note that additional DAC pathways
have been excluded (e.g., moisture swing capture) as these technologies are still in the
laboratory and demonstration phase and currently offer a potentially limited scale solution
(e.g., low concentration CO2 streams to enhance air in greenhouses).
Year DAC pathways (MtCO2)
2020 0.0028
2046 36
2080 264
2095 316.5
Table 3 - Reference scenario for DAC pathway-specific technology uptake, assuming 50% of total DAC
market for each of the pathways, based on the IEA NZE scenario total technology demand

11
11
Role of policies and public support in the normal and accelerated deployment
scenarios

Currently, DAC’s role in achieving decarbonization targets is not well recognized. Therefore,
few policies support DAC. The most significant policies and incentives are in the United States,
where tax credit is provided for direct air capture of carbon dioxide, since 2018.
Although as the technology matures its cost will inevitably fall, it is still far from being cost-
competitive. Additionally, DAC’s primary product (CO2) has limited market demand - the only
sizeable demand lies in enhanced oil recovery (EOR), which results in only a small fraction of
the total CO2 being removed. According to Meckling and Biber, this calls for policy and political
strategies beyond carbon pricing, for instance, niche market creation, building political support
and 'incentive + mandate’ strategies.
The transition to carbon removal technologies will likely differ from the solar and EV transitions
in its politics. The structure of the oil and gas industry—a sector central to the development of
DAC—differs from the utility and auto sectors. Oil and gas firms have the incentive (protecting
the cost of fossil fuels), possess the capital, the global infrastructure, and the knowledge
(chemical and mechanical processes) to develop a process like DAC (Meckling & Biber, 2021).
Government research and development (R&D) support. Governments spend around $15
billion a year on clean technology R&D, however, only a tiny fraction of this amount is
dedicated to DAC. The US government upped its game by increasing DAC R&D expenditure
from the $3 million baseline (in 2015) to $15 million (January 2021) and finally to $20 million
in March 2021. In other jurisdictions the grants for R&D available for DAC are even more
modest. Nevertheless, in 2015, a coalition of 20 countries (including the US, the UK, Japan,
China, Germany, Saudi Arabia) and the EU brought Mission Innovation to life in order to
accelerate clean-tech innovation (ICEF Roadmap, 2018)
Tax incentives. Similar tax incentives to the ones implemented in Norway to boost plug-in EV
sales could help the development and deployment of DAC. Since 2018 the FUTURE Act (45Q)
provides tax credit for CO2 storage and use in EOR by 2026 the progressive tax credit will
reach $50/tCO2 and $35/tCO2 respectively (from the respective $28/tCO2 and $17/tCO2
levels). This carbon capture specific tax credit is unique so far in the world.
Carbon price. Through emissions trading schemes and tax mechanisms, a high enough
carbon price can provide incentive to emitters to cut emissions. The current (and near-to-mid-
term) carbon price, however, is not likely to support the uptake of DAC due to the high cost of
capture. Most global carbon prices are well below the $100/tCO2 level that could help the
uptake of the technology. Nevertheless, carbon prices in the region of $20-30 can still create
incentives allowing technology developers to generate much needed project revenues.
Mandates and regulation. Government mandates can help to create markets. Early adopters
with strong support policies (e.g., subsidies) can create niche markets that help bring down
the cost of technology, as Japan and Germany have done with solar photovoltaics (Nemet,
2019). This could work as the legislation that helped create a business case for renewable
energy generation by mandating purchase or guaranteeing prioritized dispatch, like in the US
(Renewable Portfolio Standards), Germany (Renewable Energy Act), India (Renewable
Purchase Obligation), the UK (Renewables Obligation), and many other countries. Although
DAC is probably too expensive and early-stage currently to be considered for such intervention
in low-income economies, it can form part of the transition strategy in high-income ones
already. For instance, direct (e.g., EV mandates in California) and indirect deployment
mandates (e.g., low-carbon fuel standards) have been key in creating market for EVs
12
12
(Meckling and Biber, 2021). Upstream regulation (i.e., certain amount of CO2 be captured and
sequestered for every barrel of oil extracted) could help in jurisdictions that have substantial
fossil fuel reserves and it would also provide an immediately viable economic use for DAC.
Downstream regulation could be applied across almost every jurisdiction in the world. For
example, California’s Low Carbon Fuel Standard (LCFS) - adopted in 2009 with the aim to
decrease the carbon intensity of California's transportation fuel pool - mandates reductions in
the carbon intensity of transport fuels and includes credits for oil produced from DAC (Meckling
and Biber, 2021). The California LCFS and the corresponding trading mechanism, with credits
trading at an average of USD 189/tCO 29,10 can also create a business case and accelerate
deployment of DAC (and CCS) operations.
Government procurement. Government purchases can help to start and build new product
markets (e.g., low-income economies procure on average 13% of GDP, middle-income
countries procure 13.2%, and high-income countries procure 14% of GDP in goods, services
and works. Government purchases create a guaranteed market, which makes acquiring debt
capital easier for technology developers and government procurement also helps with
technology standard development, which can catalyze supply chain efficiency gains and
eventually can drive technology costs down. For example, in a similar fashion to “The Navy
Biofuel Initiative Under the Defense Production Act11”, that guaranteed drop-in biofuel
purchases, CO2-based fuels, organic chemicals or plastics procurement could be mandated.

9
https://www.iea.org/policies/11671-california-low-carbon-fuel-standard
10
https://www.neste.com/investors/market-data/lcfs-credit-price; accessed on 24/05/2021
11
https://fas.org/sgp/crs/natsec/R42568.pdf
13
13
Avoided Emissions and Emission Reduction
Potential (ERP)
Incumbent and alternative technologies
There is no incumbent technology for DAC – besides the non-existence of technological CO2
removals. Hard-to-decarbonize sectors (e.g., aviation, shipping), distributed emissions
sources (e.g., road transport), together with the Oil and Gas sector create not only a
continuous flow of CO2 for the technology – that will need it in the future to meet climate goals
due to delayed mitigation and overshooting – but can also be the primary supporters of the
technology’s development. The Oil and Gas sector can become the prime market for the
captured CO2 for its use in enhanced oil recovery or synfuels. Additionally, Direct Air Capture,
to a degree, ‘competes’ with other point-source capture technologies (e.g., BECCS) in
the downstream CO2 market. Thus, it can be argued that the incumbent technology for DAC
is early mitigation of climate change – and that provided this would occur, DAC would not be
needed. The view taken here, is that the reference scenario is one characterized by some but
insufficient mitigation – the current policy trajectories, as highlighted in several IEA and UNEP
reports – and where DAC is needed to meet the temperature goals set by the Paris agreement.
The ‘alternative’ technology in this instance is DAC with two distinct pathways, that are
considered in this methodology, namely:

 HT-DAC
 LT-DAC
The basic avoidance mechanism considered is net carbon capture (i.e., total amount of CO 2
captured minus the emissions generated during the process).

System boundary
System boundaries. Although 4 out of the 6 studies in the review consider cradle-to-grave
boundaries, 3 of these studies consider geological sequestration of the captured CO 2, which
has certain and more accurately estimated downstream GHG impacts than utilisation. 2
studies considered DAC based fuel production, applying the cradle-to-grave approach. Note
that one of the studies explores two routes for the captured CO 2, namely, sequestration and
CO2-based CH4 production/use. One study only concerned gate-to-gate (production process)
emissions with a particular focus on capture efficiency and CO 2 product purity (for potential
downstream utilisation). Regarding the processes assessed in the various life-cycle steps of
the DAC technology, unfortunately the studies showed considerable level of inconsistency.
For this review, the cradle-to-grave approach was adopted. In Figure 6 below, the main life
cycle steps that are consistently accounted for are summarized and the largest contributors to
emissions are shown in boldFigure 6. The study considering gate-to-gate energy needs and
the one considering organic chemical production with DAC CO 2 input using the cradle-to-grave
approach were omitted (further justification below).

14
14
Figure 6- System boundaries and process steps considered in the analyzed LCA literature

Functional unit. According to the ISO 14044:2006 standard, the functional unit is a “quantified
performance of a product system for use as a reference unit”. Generally, a functional unit shall
be precise and quantifiable. The kg CO2e per kg CO2 captured functional unit was chosen
for this analysis.
Impact assessment method. The impact assessment method used for calculating the GHG
emissions of DAC (Global Warming Potential (GWP) LCA indicator) is based on the IPCC-
based emissions factors expressed in terms of kg of carbon dioxide equivalent (kgCO 2eq) with
a 100-year time horizon.
Geographical presentation of studies. the reviewed studies represent predominantly
European and North American locations; although only 5 studies had any mention of
geographical locations, in total 38 sites were assessed for potential impacts related to DAC
operations (Figure 7). The choice of geo-representation affects the GHG emissions, especially
when the upstream LCA phase is considered in the system boundaries (mainly low-carbon
electricity grids were considered in the studies).

Figure 7 – Geographical representation of DAC studies


15
15
LCA Approach

The majority of LCA approaches applied in the selected studies that analyzed the two DAC
technology pathways are attributional LCA. This approach accounts for GHG emission
occurring in the life cycle phases, that fall within the considered system boundaries. The
attributional approach does not evaluate the GHG emissions displacement due to effects that
are not included in the system boundaries. In this scope the impact of the CO 2 market is not
captured in full. That is, CO2 used in enhanced oil recovery and the consequential indirect
impacts of oil production and use are not accounted for. Likewise, the social and psychological
aspects of legitimizing continued fossil fuel use are not taken into consideration. The studies
do attempt to capture downstream emissions generated by synthetic fuels using captured CO 2
as feedstock.
With regards to DACCS systems multi-functionality, emissions allocation is not a problematic
issue, and the ‘negative emissions concept’ is usually applied in the correct way, as the
captured CO2 is stored in geological formations (actual removal). The main uncertainties arise
with regards to the availability carbon storage sites only.
However, if the captured CO2 is utilized, due to the wide variety of uses, allocating impacts
can become problematic. For example, considering EOR, fuel and chemicals production are
especially important as in these processes' combustion and other End-of-Life emissions are
very relevant. The energy content allocation approach is the most used in the LCA literature
(Müller et al., 2020). It allocates the total GHG emissions generated from the entire life cycle
of the products (including combustion) proportionally to their total energy content.

Reference scenario
In this instance, the incumbent technology case is a world without CO2 capture and insufficient
mitigation measures. The avoidance mechanism considered is ‘net carbon capture with CO2
lifecycle emissions’. This takes an attributional life cycle approach which is used to calculate
the GHG footprint and emissions reduction potential (ERP) for the technology (actual capture).
To calculate the ERP, it is necessary to define a reference scenario for the adoption of the
DAC technology. For this purpose, the scenarios defined in the “Market Uptake” section are
used. For market uptake projections the IEA Net Zero (NZE) model was considered as the
reference scenario that is accelerated via additional investment.

GHG emission rates


Please note, the default value throughout the document and in all subsequent estimations
(e.g., carbon abatement factors, ERP) is the average value, however, the minimum and
maximum values are also given to demonstrate the inherent uncertainty of any of the
calculated figures.

Incumbent

As explained previously, there is no ‘real’ incumbent technology to measure DAC against,


which poses some methodological challenge. In principle, the current situation where
technological removals are negligible and emissions are high, is taken as the incumbent case.
Net carbon capture considering CO2 life cycle emissions are compared to a situation with no
net carbon capture.

16
16
Alternative technologies

Permanent storage might lead to negative emissions already in many countries today,
however, the energy mix’s emissions intensity is crucial; according to Deutz et al. (2020), most
system configurations become climate positive when the emissions intensity of the grid is on
par or lower than 256.2gCO2/kWh. The life cycle impacts in all cases of sequestration show
better results, i.e., smaller (or negative) GHG footprint. Sequestration is also considered safe,
leakage is negligible (potential leakage:<2% of sequestered CO2), however the presence of
suitable storage sites and minimal transport is key in minimizing the impacts (Kelemen et al.,
2019). DAC as renewable carbon feedstock for fuels could also reach carbon neutrality over
the entire life cycle but relies on low-carbon energy supply, hence large-scale DAC operations
must be located close to renewable energy sources (Deutz et al., 2020).
The results for HT-DAC must be heavily caveated as only one study, de Jonge et al. (2019)
explored the life cycle impacts of this technology option. Besides that, the only other literature
available is a partial LCA assessment by Keith et al. (2018), and only carbon storage was
explored for the pathway.
The LT-DAC system had a significantly better coverage and therefore the LCA emissions
results are more reliable. Of the 39 studied configurations (variables: heat and electricity
source, geographical locations, captured CO 2 utilization) 35 explored the impacts of geological
sequestration as final use for the captured CO 2. 4 cases considered fuel production (and
combustion), namely, synthetic methane (power-to-gas) and synthetic diesel production (with
Fischer–Tropsch synthesis (FTS), however, these were excluded from the assessment, as in
the IEA models it is assumed that around 95% of total CO2 captured in 2050 is stored in
permanent geological storage. Please see the overall CO2 impacts in Table 4.
HT-DAC - storage LT-DAC - storage
(tCO2/tCO2 (tCO2/tCO2
captured) captured)
tCO2 emissions, average 0.453 0.141
tCO2 emissions, min 0.080 -0.951
tCO2 emissions, max 0.900 0.914
Table 4- LCA emissions intensities of the technology pathways

The direct climate impacts of the DAC systems can be greatly reduced if RE is used to provide
the system’s energy needs, however, the upstream impacts of building solar PV panels or
commissioning a solar PV farm (exclusively for the DAC plant) can add to the overall footprint
of the operations and can have significant impact on the initial capital cost of the plant as well.
Note that recycling of metals has a negligible impact on the overall GHG footprint
(9gCO2/kgCO2 captured). Amine production also has a relatively small climate effect, namely
0.9-23 gCO2/kgCO2 captured, across 66 different amine-based sorbents (Deutz et al., 2020).
Emissions related to the transportation and storage of the captured CO 2 can make up as much
as 5-15% of the total carbon footprint of the process dependent on the DAC plant's location
from the location that is designated for CO 2 storage.
As mentioned before, the most important factors, besides the end use, for the DAC technology
in terms of climate impact are the emissions intensity factors of the energy used in the process
and the geographical location of the plant (for electricity mix, infrastructure and ready-to-use
storage sites).

17
17
Direct and indirect emissions: carbon storage and utilization

The final product (utilization) defines how energy intensive the system is and defines how
much energy is used by the DAC plant (direct impacts). For example, if the intended use of
the captured CO2 is microalgae production, CO2 purity can be as low as 5% (usually between
5-35%) and the process would require between 216-440 kJ/mol CO2 captured. However, if the
aim is to produce a highly compressed, pure CO 2 stream for sequestration or beverage
production (99% purity) the process would require between 635-1501 kJ/mol CO2 energy,
which can come with substantial emission if fossil energy is utilized (Table 5). However, using
renewable energy sources can significantly reduce the emissions occurring during the capture
process, and a ~90% capture efficiency can halve the energy required for capture. Moreover,
the final utilization (and the required CO2 stream purity) has a significant impact on capture
costs as well (Wilcox et al., 2017).
As mentioned above the captured CO 2 can be sequestered or utilized. Captured CO2 can be
used in enhanced oil recovery (EOR), in the production of fuels, chemicals, building materials
(improving concrete strength, creating a carbon neutral concrete alternative). Moreover, CO2
can be used to produce carbonated beverages, and it can be used to enhance the productivity
of algae farms and to enrich the air in greenhouses. As it can be seen in the above examples,
many of the utilization pathways on the one hand require a highly pure CO 2 feedstock, and on
the other, the CO2 is quickly re-released to the atmosphere.
The different uses of the captured CO2 can have very different potential environmental impacts
and can create different business opportunities, especially if policy or other incentives are in
play. On the one hand, when CO 2 is geologically stored, it is permanently removed from the
atmosphere, resulting in negative emissions. On the other hand, most large-scale
opportunities (gigaton-level) to use the captured CO 2 result in its re-release into the
atmosphere (e.g., when synthetic fuel is burned). Although this re-release will not result in
negative emissions, it could still generate climate benefits, for example if conventional fossil
fuels are replaced with synthetic fuels (that have smaller life cycle carbon footprints than their
‘dirtier’ counterparts).

Low purity (5-35%) Medium (50%+) High purity (90%+)


(kJ/mol CO2) purity (kJ/mol CO2) (kJ/mol CO2)
50% capture
216-440 493-635 635-1501
efficiency
90% capture
130-274 308-399 399+
efficiency
Potential uses Microalgae production EOR, ECBM Beverage production,
sequestration, EOR
Table 5 - Energy requirement of CO2 capture at different efficiency and purity levels (Wilcox, 2017)

1) Low purity (5-35%) utilization.


By far the cheapest, least energy intensive pathway. In the near-term it would be crucial to
identify opportunities that can use low-concentration CO2 as feedstock as the bottom-line cost
of CO2 capture could be significantly lower. Mineral carbonation, where magnesium or
calcium oxides react with CO2 to form carbonates, in theory can be a cheap way to store CO2
for long periods with low-energy requirement, additionally, impurities (e.g., NOx) do not
interfere with the carbonation reaction (Lackner et al., 1995). Biofuels through microalgae
production, has a few advantages, such as it does not require a pure CO2 stream as an input
and microalgae cultivation does not represent a threat to food markets. Nevertheless, the high
production cost of transforming microalgae biomass to useful biofuels products might limit this

18
18
opportunity (Cuéllar-Franca, 2015). The dietary supplement market simply represents a
negligibly small market for CO2 products (Wilcox, 2017).
2) ‘Medium’ purity (50%+) utilization
Enhanced Oil Recovery (EOR) and Enhanced Coal Bed Methane (ECBM) recovery. The
most commercially advanced opportunity is EOR, and although dilute CO 2 can be used in the
process, recovery may decrease with dilution. Dilute CO2 for EOR, including gas mixtures of
N2 and CO2,, recovers only 80%–85% as much oil as pure CO2 at best. Although EOR will not
result in negative emissions, according to Occidental, if the EOR process is linked to DAC
(using renewable energy sources) the amount of CO2 released during the lifetime of the fossil
fuel products12 might equal the amount of CO2 sequestered in the EOR process. EOR and
ECBM represent probably the largest market for captured CO 2. ECBM using CO2 has an
estimated 140 Gt of CO2 of storage following methane recovery in China alone (Brickenhoff,
2011). There are currently around 166 projects using CO2 in EOR, producing around 0.55
million barrels per day of oil. The size of this market along with the existing government
subsidies and incentive schemes (e.g., 45Q in the US) represent a sizeable opportunity for
captured CO2 utilization.
3) High purity (-99.9%) utilization
Food and drink industry, and pharmaceuticals. These markets are restricted by not only
size but also to sources that produce high purity CO 2 streams, potentially making the capturing
process too expensive to be viable without any additional policy incentive. Similarly to
synthetic fuels, these uses of captured CO2 offer only a temporary storage and CO2 is then
quickly re-released to the atmosphere.
Conversion (fuels). Although highly pure CO2 is preferred in fuel production, as long as the
gas in which CO2 is diluted is not competing in the chemical reaction (such as the highly inert
nitrogen gas), lower purities can work as well (Kirchofer, 2012). The drawbacks of CO2 based
synthetic fuels is that the conversion is energy intensive, and the CO 2 is released into the
atmosphere before the benefit of the capture can be realized (the fuels offer a limited, 6-
months storage period for CO2). Capturing CO2 via DAC and using CO2 as a feedstock in FT
fuel production is not a net negative emissions pathway but the climate impacts of the resulting
fuel can be lower than conventional fuel production (common emissions factor for diesel fuel
at 75 gCO2e per MJ) (Liu et al., 2020). Note, the results in Figure 8 assume that in the DAC-
FTS process only renewable energy is used. If this assumption is not met, the climate
outcomes of the DAC-FT fuel can be much worse than conventional fuels. The breakeven
point of the electricity supply is 139 gCO 2e/kWh. If the emissions intensity of the supply
exceeds this value, the climate impacts of the DAC synthetic diesel will be worse than that of
conventional diesel.

12
https://www.oxy.com/Sustainability/overview/SiteAssets/Pages/Social-Responsibility-at-
Oxy/Assets/Occidental-Climate-Report-2019.pdf
19
19
Figure 8 – Comparison of carbon intensity of different diesel production methods: conventional, biodiesel,
CO2 to fuels, and the combined direct air capture (DAC) and Fischer–Tropsch synthesis (FTS) process (Liu
et al., 2020)

Conversion (chemicals). As long as the energy used in the DAC and water electrolysis
processes (e.g., wind electricity and waste heat), the overall lifecycle emissions of organic
chemicals can be significantly lower than their fossil-based counterparts (Rosental et al. 2020),
given that the captured CO2 amounts are netted from the overall carbon footprint of the
chemicals (Table 6).
DAC process emissions LCA emissions using DAC LCA emissions using fossil
reference products
Methanol 265.27 kgCO2e/1t Methanol 1375 kgCO2e /1t Methanol 2002 kgCO2e /1t Methanol
Olefins 763.58 kgCO2e /1t Olefin 3142 kgCO2e /1t Olefin 4591 kgCO2e /1t Olefin
BTX31 725.20 kgCO2e /1t BTX 3352 kgCO2e /1t BTX 4648 kgCO2e /1t BTX
Table 6 - GHG emissions during the DAC process and product life cycles for organic chemicals using
captured CO2 as feedstock vs their fossil reference counterparts (Rosental et al., 2020)

Sequestration/storage. Probably the only pathway that is fully climate positive, resulting in
net negative emissions. Although some leakage might occur at storage sites (up to 2% of the
total amount of CO2 sequestered), it is likely that storage sites will retain the stored CO 2 for
over 1,000 years (Kelemen et al., 2019). However, without additional incentives (such as the
45Q tax credit in the US) the potential market remains small, currently in the region of 40
million metric tons of CO2.

GHG impact assessment method

The common impact assessment method used for the GHG emissions calculation is the
Global Warming Potential (GWP) with a time horizon of 100 years (IPPC, 2019), expressed in
carbon dioxide equivalent (CO2eq).
The life cycle GHG values calculated for DAC are deducted from one metric ton of atmospheric
CO2 (technical flow, that is anthropogenic in origin). The calculation included emissions
generated during the operational processes and the considered upstream and downstream

20
20
life cycle phases, and emissions embedded in the material and energy flows as inputs to the
system.

Carbon Abatement factor


The Carbon Abatement Factor (CAF) was calculated, for each technology pathway, in
accordance with the expression13.

𝐶𝑎𝑟𝑏𝑜𝑛 𝑎𝑏𝑎𝑡𝑒𝑚𝑒𝑛𝑡 𝑓𝑎𝑐𝑡𝑜𝑟𝑦


= 𝐸𝑚𝑖𝑠𝑠𝑖𝑜𝑛𝑠 𝑜𝑓 𝑖𝑛𝑐𝑢𝑚𝑏𝑒𝑛𝑡 𝑡𝑒𝑐ℎ𝑛𝑜𝑙𝑜𝑔𝑦 𝑝𝑒𝑟 𝑓𝑢𝑛𝑐𝑡𝑖𝑜𝑛𝑎𝑙 𝑢𝑛𝑖𝑡𝑦
− 𝐸𝑚𝑖𝑠𝑠𝑖𝑜𝑛𝑠 𝑜𝑓 𝑎𝑙𝑡𝑒𝑟𝑛𝑎𝑡𝑖𝑣𝑒 𝑡𝑒𝑐ℎ𝑛𝑜𝑙𝑜𝑔𝑦 𝑝𝑒𝑟 𝑓𝑢𝑛𝑐𝑡𝑖𝑜𝑛𝑎𝑙 𝑢𝑛𝑖𝑡𝑦
Where:
incumbent technology considered was one metric ton of atmospheric CO2
y is year.

In order to calculate the Carbon Abatement Factor (CAF) and to make it compatible with the
principles of avoided emissions the incumbent technology case considered was a world
without DAC where carbon removals would not happen. Thus, the CAF is considered as the
net flow of CO2 emissions of the DAC technology, including the potential negative net flow of
CO2 – this is, if the DAC process emits 0.1 tCO 2 for each metric ton of CO2 captured the net
flow will be 0.9 tCO2. CAFs for the different pathways are presented in Table 7 and Figure 9.

CAF - GHG (tCO2/tCO2)


Average Best Worst
HT-DAC (storage) 0.547 0.920 0.100
LT-DAC (storage) 0.859 1.951 0.086
Table 7- Carbon Abatement Factor for the DAC pathways

Carbon abatement factor (aver, min, max)


2.5
GHG (MtCO2eq/MtCO2eq captured)

1.5
Avr
Min
1
Max

0.5

0
HT-DAC LT-DAC

13
According to the above, if the comparative impact is zero, the assessed and reference products emit
the same amount of GHGs over their life cycles. If the comparative impact is negative, the assessed
product creates more environmental load over its life cycle compared to the reference product. If the
comparative impact is positive, the assessed product emits less compared to the reference product
over its life cycle (Russell, 2018)

21
21
Figure 9- Carbon Abatement Factor for the DAC pathways

As can be seen, depending on the specific technology pathway, energy source (for electricity
or heat), and designated downstream utilisation might not always be beneficial in terms of
GHG emissions. In the case of the 2 technology pathways, the CAF is considered constant
throughout time14.

For this methodology, the default value is the average value. However, we will refer to the Min
and Max (Best/Worst) values as ways of characterizing the inherent uncertainty of final figures.

Emission Reduction Potential


With regards to the carbon-dioxide removal (CDR) technologies, it is important to distinguish
between negative emissions (e.g., permanent CO2 removal) and avoided emissions (e.g., CO2
utilization as feedstock). Real negative emissions can only correspond to permanent removal
from the atmosphere, which DAC pathways can deliver, if the captured CO2 is stored
permanently (Terlouw et al., 2021).
Annual ERP of any given technology pathway can be calculated using the CAF, as shown
below,

𝑛
𝐸𝑠𝑡𝑖𝑚𝑎𝑡𝑒𝑑 𝑉𝑜𝑙𝑢𝑚𝑒 𝑆𝑐𝑒𝑛𝑎𝑟𝑖𝑜𝑦 ∗ 𝐸𝑠𝑡𝑖𝑚𝑎𝑡𝑒𝑑 𝐶𝑎𝑟𝑏𝑜𝑛 𝐴𝑏𝑎𝑡𝑒𝑚𝑒𝑛𝑡 𝐹𝑎𝑐𝑡𝑜𝑟𝑦
𝐸𝑅𝑃𝑡𝑒𝑐ℎ = ∑ ( )
(1 + 𝑟)𝑦
𝑦=1

Where
Estimated Volume Scenario: is the volume of product or activity for a given technology
in a given uptake scenario
Estimated Carbon Abatement Factor: the CAF for a given year, calculated as above
y: is a given year
r: is an avoided emissions discount rate

Annual ERP using the CAF in Figure 10.

14
The Carbon Abatement Factor (CAF) is taken as a constant as it is considered the conservative
approach. The CAF might vary not only regionally (e.g., according to the local electricity grid emission
factor) but also over time (i.e., the technology might improve, and the CAF might become higher) leading
to very different outcomes, using the same CAF over time avoids overestimating the impact of the
technology.
22
22
ERP for each DAC pathway using CAF (50%
market share per technology)
40
35
30
ERP (MtCO2)

25
20
15
10
5
0
2020 2025 2030 2035 2040 2045 2050 2055

HT-DAC LT-DAC HT-DAC (discounted) LT-DAC (discounted)

Figure 10 - Annual ERP for each DAC pathway using potential capture projections and CAF in the
reference scenario (for discounted ERP a 3% discount rate was applied)

Avoided emissions
Avoided emission at fund level for a given year y, can be determined using the equation

𝐴𝑣𝑜𝑖𝑑𝑒𝑑 𝐸𝑚𝑖𝑠𝑠𝑖𝑜𝑛𝑠𝑓𝑢𝑛𝑑, 𝑦
𝑛

= ∑ 𝐶𝑎𝑝𝑡𝑢𝑟𝑒𝑑 𝑉𝑜𝑙𝑢𝑚𝑒𝑠𝑝,𝑦 ∗ 𝑀𝑒𝑎𝑠𝑢𝑟𝑒𝑑 𝐶𝑎𝑟𝑏𝑜𝑛 𝐴𝑏𝑎𝑡𝑒𝑚𝑒𝑛𝑡 𝐹𝑎𝑐𝑡𝑜𝑟𝑝,𝑦


𝑝=1
Where:
𝑛: is the number of DAC projects in the portfolio
𝐶𝑎𝑝𝑡𝑢𝑟𝑒𝑑 𝑉𝑜𝑙𝑢𝑚𝑒𝑠𝑝,𝑦 : are the captured volumes of CO2 for project p in year y

𝑀𝑒𝑎𝑠𝑢𝑟𝑒𝑑 𝐶𝑎𝑟𝑏𝑜𝑛 𝐴𝑏𝑎𝑡𝑒𝑚𝑒𝑛𝑡 𝐹𝑎𝑐𝑡𝑜𝑟𝑝,𝑦 : is the Carbon Abatement Factor calculated


for that project and year

Uncertainty and challenges

Uncertainty related to methodology

There are several methodological options that can introduce uncertainty into the final results,
the most prominent being, the allocation procedure. The next paragraphs discuss different
sources of uncertainty and their relevance.
Allocation procedure
With regards to DACCS systems multi-functionality and emissions allocations is not a relevant
issue and the ‘negative emissions concept’ is usually applied in the correct way (Terlouw et
al, 2021). The main uncertainties arise with regards to the availability of carbon storage sites
only.

23
23
However, if the captured CO2 is utilized, due to the wide variety of uses allocating the impacts
can become problematic. For example, considering EOR, fuel and chemicals production are
especially important as in these processes, combustion, and other End-of-Life emissions are
very relevant. The energy content allocation approach is the most used in the LCA literature
(Müller et al., 2020). It allocates the total GHG emissions generated from the entire life cycle
of the products (including combustion) proportionally to their total energy content.

Captured CO2 is often treated intuitively as a “consumed” emission


(or elementary flow). However, treating captured CO2 as a negative emission is usually
incorrect and if used as a chemical feedstock for CO2 utilization captured CO2 must be treated
like any other feedstock (Müller et al., 2020). The principles on how to account for CO 2
throughout the entire life cycle can be seen in Figure 11, adapted from the analogous CCU
technology (the only difference being the source of CO 2).

Figure 11– Schematic life cycle of CCU technologies span from the CO 2 source, supply other feedstocks
and energy to the end-of-life treatment. In all life cycle stages, environmental impacts should be
considered (von der Assen et al., 2013).

The overall climate impact and the sustainability of the technology can be substantially
affected by the actual DAC processes (e.g., type of energy input and system configuration)
and the use of CO2-based products. With regards to DAC-captured CO2 utilization LCA
literature is scarce. CO2 utilization does not necessarily contribute to addressing climate
change, and careful analysis is essential to determine its overall impact. Identifying the
counterfactual - what would have happened without CO 2 utilization - is important but is often
particularly challenging (Hepburn et al., 2019).

Probability of success
The Avoided Emissions Framework considers that when applying its approach to future
scenarios “then the probability of the solution delivering the expected benefits should be
included (probability of successful development of the solution), as well as the probability of
the solution being adopted at scale (probability of adoption)”.
In the case of the ECT Framework, the probability of success is reflected by using the efficacy
of investment such that an investment of $500 million over two years yields an efficacy of 70%,
this is $350 million.

Uncertainty related to GHG data


An inherent issue with the interpretation of carbon dioxide removal LCA studies is that
emissions avoided can be easily misinterpreted as negative emissions (carbon removal from
the atmosphere). Emissions from the full DAC process can be potentially negative, if CO 2 is
captured and sequestered or permanently stored in a product (e.g., concrete) and if overall
24
24
lifecycle of GHG emissions of a product is lower than the amount of CO2 fixated (Müller et al.
2020). The DAC process can be carbon neutral if the amount of atmospheric CO 2 capture and
fixation is equal to other fossil emissions over the life cycle. In all other cases, DAC
technologies have positive CO2 emissions over the life cycle, nevertheless, emissions can be
lower than for competing conventional processes.
The technology itself is new (Technology Readiness Level 6), and enough large-scale
demonstrations have not yet taken place to identify learning and cost saving opportunities on
economic scale. Lack of dedicated, accessible, and transparent research also carries
substantial risk of misinterpreting avoided emissions or economic data.
Most DAC studies are assuming the use of low-carbon energy sources (PV, wind) or
geographical locations with relatively low grid average emissions factors (e.g., large volumes
of hydro power in BC, low-carbon grids of Iceland or Switzerland). System configuration can
also have huge influence on the final GHG impact and initial capital costs (e.g., grid electricity
and waste heat versus solar/wind and waste heat). The most promising outcomes, and
therefore what technology deployment depends upon, are reliable renewable energy (RE)
pathways and future electricity mix changes that assume even higher rate of RE penetration.
As such, DAC relies on policy instruments that facilitate RE generation.
Land and location are concerns, as DAC operations require buildings (land area) and
infrastructure that serve those buildings to ensure plants operate effectively (e.g.,
transportation cost reductions, water requirements, access to clean energy, etc.).
Furthermore, atmospheric conditions are also an important factor in selecting the location
(humidity, pressure, density) and these factors can have a substantial impact on the water
consumption (process losses), and heat energy requirement of the technology.
As to carbon storage and use, the technology requires a stable carbon economy, a market for
carbon dioxide. Beyond carbon storage and EOR, other large-scale markets must be explored.
Although CO2 is a feedstock in many processes and consumers exist, they only represent a
megaton-level market (e.g., food and beverage) or the carbon storage capacity of the products
can prove insignificant (e.g., plastics, synthetic fuels).

Challenges related to secondary enabling and rebound effects

The most substantial issue with regards to rebound effects might arise from the market
dynamics for captured CO2. The most developed markets that could take up large amounts of
CO2 feedstock are the EOR and synthetic fuel markets. By the nature of these markets the
captured CO2 could be soon re-released to the atmosphere (by means of combustion), even
before the benefits of capture could realize themselves.
DAC could also (1) postpone the timing of mitigation as it can be deployed later in the century
at a significant scale (reducing short-term costs), (2) DAC reduces the marginal and total
abatement costs and, (3) DAC allows for a prolonged use of oil, with a positive welfare impact
for energy exporting countries (Chen & Tavoni, 2013). Direct air capture maintains the value
of oil in the ground despite carbon constraints by allowing the continued use of liquid fuels in
the transportation sector (Lackner, 2009).
Additionally, the availability of DAC technologies reduces the stranding of existing coal and
gas based conventional power plants and delays any stranding further into the future. The
carbon budget to meet the climate goal of limiting end-of-century warming to 1.5–2°C would
require abating 28–33% of 564 Gt CO2 (the total committed CO2 emissions from the existing

25
25
power plants) vs. a 46–57% reduction in the scenario without direct air capture and carbon
storage technologies (Pradhan et al., 2021).
We would like to note that rebound effects go beyond the boundaries and scope of the model
and attributional approaches used. There is an overall lack of consequential LCA studies for
many of the low-carbon technologies needed to meet the Paris goals. The goal of the ECT
Framework is to estimate the investment required to accelerate the development of low carbon
technologies and the behaviour of the final users is not part of the models used in this
methodology. For this reason, it will be particularly important to be aware of the potential
rebound effects and monitor if they are occurring or not.

26
26
Reductions in Green Premium
The Green Premium is defined simply as the cost of DAC (for either of the two technological
pathways).
𝐺𝑟𝑒𝑒𝑛 𝑃𝑟𝑒𝑚𝑖𝑢𝑚 = 𝐶𝑜𝑠𝑡 𝑜𝑓 𝐷𝐴𝐶
The reductions in Green Premium are calculated as a % decline of an initial Green Premium
and the Green Premium at time t, as Per equation below.
𝐺𝑟𝑒𝑒𝑛 𝑃𝑟𝑒𝑚𝑖𝑢𝑚𝑡
𝑅𝑒𝑑𝑢𝑐𝑡𝑖𝑜𝑛 𝑖𝑛 𝐺𝑟𝑒𝑒𝑛 𝑃𝑟𝑒𝑚𝑖𝑢𝑚 = 1 −
𝐺𝑟𝑒𝑒𝑛 𝑃𝑟𝑒𝑚𝑖𝑢𝑚𝑡=0
Market prices will vary dependent on costs (namely raw materials and energy costs), on
factors linked with market power and competition, as well as varying with time and geography.
It might be needed to calculate Green Premiums with a mix of market prices and production
costs.
Obtaining market price information at this stage is limited to the price listing of commercially
operating DAC developers. Regarding market price drivers (Capex, operations, and
maintenance costs) see more details in the ‘Cost models section’. Currently only Climeworks
has an openly accessible pricing structure, prices are below.

 LT-DAC price: 850-895 $/tCO2 (dependent on subscription type)


These values can be used for future reference in calculating Green Premium reductions based
on price data, although further analysis and data collection is recommended looking at both
inter-annual variation of Green Premiums as well as regional variation, which will be done as
soon as data is available.

27
27
Learning curves
The academic literature seems to be more aligned with regards to learning rates of the
technology. Nemet and Brandt (2012) derive their learning rates from previous work on
carbon capture at power plants, the closest analogy for air capture. They consider learning
rates of 0.101 for capital costs, 0.135 for energy costs and 0.135 for operational and
maintenance cost. Estimates for the energy needs – and by extension energy and
operational costs – of DAC processes are a point of contention. Theoretically, capturing CO2
from ambient air requires very little energy, however, this energy requirement in practice is
much higher. Moreover, the different technologies require different amounts of energy based
on the combination of processes used, and they require different qualities of energy as well
(i.e., low and high temperature heat) (Broehm et al., 2015).
Furthermore, near term capital costs are also somewhat uncertain; for instance, factory-type
designs (shipping container designs excluded) show a wide range between $300 million and
$3 billion, with a middle estimation of $1.6 billion. These costs can be expected to fall due to
the economies of scale and optimization of the systems and drop to between $20 and $200
million on the long term (Broehm et al., 2015), indicating a learning rate between 10% and
15%. The most likely candidate to deploy faster initially among all different approaches, is the
most developed DAC system, the aqueous solution design. This system has shown significant
technological improvement over the past years and will continue to follow the pattern which in
the long term will bring capital and operational expenditures down (Broehm et al., 2015).
Fasihi et al. (2019), considers a conservative (10%) and a high (15%) learning rate in
exploring the potential technology diffusion for DAC. Although, the 15% learning rate
matches the technology specific characteristics of DAC systems better, it assumes the
effective and timely execution of the Paris Agreement (including carbon pricing and
accelerated carbon removal). On the other hand, the conservative scenario with a 10%
learning rate for DAC systems (based on sulphur removal systems of large-scale coal plants)
seem to be more aligned with current trends, as it assumes the delayed execution of the Paris
Agreement. This lower learning rate and slower cost reduction leads to a 50% reduction in the
cumulative potential of CO2 captured by DAC systems, however, the roughly 7.8 GtCO 2 annual
capture is still in line with the projections of IEA’s SDS scenario.
In summary, research seems to agree that a 10% learning rate is expected and realistic for
the technology, however, with the right policy and financial incentives, a 15% learning rate is
likely (Table 8).
DAC Pathway Learning rate and comments
LT-DAC 10% is a highly probable learning rate for the technology, the 15% learning
HT-DAC rate requires additional policy support (Fasihi et al., 2019)
Table 8 - Learning rate used in the modelling of learning curves for each SAF technology pathway

This means that the necessary investments in the reference scenario can be modelled by
considering the increase in capacity year-on-year derived from the reference scenario and
which should translate in an additional cumulative installed capacity.
The equation describing the relationship between production quantity (Q) and cost (C) follows
an exponential decline and is known as the learning curve. The governing equation for the
learning curve is:

𝐶 = 𝑎𝑄−𝜆

28
28
The learning rate, λ, in the learning curve is the same as that from the diffusion equation, and
thus connects cost declines with market adoption. The learning curve equation constant, a,
can be interpreted as a “first unit cost” that is solved for with known pairs of cost and quantity.

Cost models

It is difficult to find a specific cost for DAC because each method has wide variation in plant
size and setup, sorbent regeneration (OPEX) and capital costs. Proposed plant sizes vary
between small modular solutions to large chemical-plant structures. This offers a wide range
of potential capital costs, between $300 million and $3 billion, that gives a reasonable middle
estimate of $1.6 billion for near-term one megaton DAC plant upfront capital costs (Broehm,
2015). The bare equipment and total initial capital costs (that include installation, labour,
engineering, and materials costs) are shown in Table 9 in 2020 USD. Where total
installation/capital costs were not available, optimistic and pessimistic costs were estimated
using the methodology put forth by Socolow et al. (2011), where for optimistic costs a 4.5-fold
and for pessimistic costs a 6-fold multiplier was applied to the total equipment costs. These
factors were suggested based on observation on post-combustion carbon capture (PCC)
operations. At the technology’s current level of technological maturity, the pessimistic values
are likely to be more realistic.
The total capital costs (estimates) account for a wide range of construction and other costs
including engineering, owner’s costs, site preparation, foundations and installation, piping,
tanks, control systems, buildings, utilities, start-up expenses and contingencies. However, the
below estimates have an accuracy of +/-50%, and experience shows that in the case of early-
stage technologies the +50% is more likely (Socolow et al., 2011). Typical CAPEX breakdown
by the National Academies of Sciences, Engineering, and Medicine (NASEM) can be seen in
Figure 12.

Figure 12 – Typical Capital cost breakdown for DAC technologies (NASEM, 2019)

29
29
Table 9 - Various DAC system configurations, initial capital costs (equipment and total installation costs).
*N.B. Keith et al., provided measured, not estimated costs of a plant being commissioned currently, where a 10k tCO2 plant is upscaled to a megaton scale plant.

30
30
Operational cost. Such estimates are equally hard to gather. It can be observed that the
annual operational costs are systematically lower than the annualized capital costs. Much of
the operating costs depend on the source and cost of energy. Renewable energies increase
the initial capital costs, however as a result, operational cost will be surely lower (Broehm et
al., 2015). The American Physical Society’s APS report conservatively caps the operational
costs at $200/tCO2 (with a lower estimate of $170/tCO 2), this value seems to be in line with
most research and industry reported values. The lower end of the OPEX cost estimates comes
from the National Academies of Sciences, Engineering, and Medicine (2019). Their research
suggests that using cheap natural gas in a liquid solvent DAC facility, operational expenditure
can be as little as $66/tCO2, but even for solid sorbents the OPEX costs are capped at
$94/tCO2. These latter values have also been confirmed by Fasihi et al. (2019).
Levelized cost of capture. The final cost of capture is a combination of the annualized cost
of capital costs and operational (and maintenance) costs. We use the model described by
Sokolow (2011) and Thunder Said Energy (2019): the discounted cash flow (DCF) method.
This method is in line with the methodology used in the ECT Framework, and it is used to
calculate the Minimum Selling Price (MSP) of one metric ton of captured CO2, or in other words
the annualized cost of CO2 capture.

 The annualized total cost of CO2 capture is the sum of depreciation, return on
investment, and all fixed and variable operating costs, using a modest 7% capital cost.
 Components of annualized total cost are divided by metric tons of CO2 captured during
the year to arrive at components of the cost per metric ton of CO2 captured.
𝑛
𝐶𝐹𝑖
𝐷𝐶𝐹 = ∑
(1 + 𝑟)𝑖
𝑖=1

Where:
𝐷𝐶𝐹: Discounted Cash-flow;
𝐶𝐹𝑖 : net cash flow in year I;

𝑟 : is the discount rate for the value of future cash;


𝑛: is the lifetime of the project;
The net cash flow is calculated using several variables that might be specific to certain
technology pathways. For the sake of consistency and comparability, where the variables are
the same (e.g., cost of H2 gas) the same prices are used or, if not, the differences are justified.
𝐶𝐹𝑖 = 𝑅𝑖 − (𝐼𝑖 + 𝑀𝑖 + 𝐸𝑖 )
Where:
𝐶𝐹𝑖 : Net cash flow in year i;
𝑅𝑖 : Revenues in year i;
𝐼𝑖 : Investments costs in year i;
𝑀𝑖 : Maintenance, labour and chemical costs in year i;

𝐸𝑖 : Energy (power and fuel) costs in year i;

31
31
Error! Reference source not found. and Table 11 show the main parameters that need to
be set to fully specify the costs model for the technology pathways.
Parameter Value Comments
𝑛 (years) 20 / 30 Sokolow et al. (2011) assumes a 20-, whilst Thunder Said Energy
(2019) a 30-year lifetime for the projects.
Fasihi (2019), in his techno-economic analysis, sets the lifetime of
LT-DAC technology to 20 years, which he expands to 25 and 30
years in 2030 and beyond, respectively, according to the long-term
estimations for generic DAC plants in literature. The lifetime of HT-
DAC is set to 25 years in 2020 according to Keith et al. (2018), and
Fasihi (2019) expands it to 30 years beyond 2030.
𝑟 (%) TBD Sokolow et al. (2011) sets the discount rate to 7%, based on
or observations made on post-combustion capture plants.
7%
Depreciation Straight Sokolow et al. (2011) and Thunder Said Energy (2019) both apply
schedule line a straight-line depreciation. This default assumption is used
throughout the techno-economic analyses cited.
Depreciation 15 Aligned with Thunder Said Energy (2019), as this reflects the
period (years) standard accounting practices for new facilities in techno-
economic assessments.
Inflation rate TBD Specific to each region where the analysis is being done.
(%)
Table 10 - Key parameters to be considered in cost and revenue calculations

For the sake of this methodology the following cost categories should be considered in the
calculation of costs and revenues.
Cost Comments
𝐼𝑖 Sokolow et al. (2011) suggests calculating total capital costs as a product of the
total major equipment purchase cost (bare equipment costs) and a capital
multiplying factor.
For the DAC system, an optimistic capital cost estimate would use the
multiplying factor of 4.5, the same as the factor that was observed in the post-
combustion capture case. A higher multiplying factor, such as 6, could be
justified to generate a potentially more realistic capital cost reflecting the novelty
of the major DAC unit operations.
𝑀𝑖 Maintenance and labor costs are incurred to keep their assets in good working
condition, comprise maintenance, insurance and overhead and can be set as a
3.7% and 4% of CAPEX for HT-DAC and LT-DAC, respectively (Fasihi et al.,
2019).
𝑀𝑖 = 𝑀% ∗ 𝐶𝐴𝑃𝐸𝑋

𝑀%: Maintenance costs


𝐸𝑖 Energy (fuel and utility prices) and consumable (chemical and water) prices and
costs are independent of the capital costs.
The different technologies and designated utilisation purities come with different
capture related energy needs, however, the efficiency of the process (i.e.,
energy need) can be simplified considering simply total production and overall
efficiency of the process.
𝑅𝑖 Revenues for year i calculated from the sales of the product output (i.e., CO 2)
and the estimate of their market prices. In this case, revenues consider the
“Minimum Selling Price (MSP.
Table 11 - Revenue calculation and main cost categories and their variables for each technology pathway

32
32
In Table 14 the levelized costs of production ($/tCO2) can be seen in 2020 USD. The cost
ranges (average and ceiling) were established using information from the available literature.
Averages were calculated from the reported values (2020 $) and gauged against Fasihi’s
extensive work on levelized costs (N.B. for LT-DAC the calculated averages were ~30% higher
than the price calculated by Fasihi (2019), however, expert interview with ClimeWorks
suggested using the higher average is justified). The price ceiling was taken from Sokolow
(2011) for HT-DAC (pessimistic scenario) and Kiani (2020) for LT-DAC from (standard
chemical engineering design using a variety of economic factors).
Green Premium for the technology (not having a counterfactual) cannot be calculated, the
levelized costs presented can be taken as “baseline figures” against which the future cost and
Green Premium reductions can be measured. For reference, levelized costs of point-source
capture (PCC) (Sokolow et al., 2011) are also shown in Table 12.

Levelized costs of capture


Technology HT-DAC ($/tCO2) LT-DAC ($/tCO2) PCC ($/tCO2)
Price ceiling 550 730
Average price 243 303 62
Green Premium
against point
source capture 181 241 n.a
(PCC) (average
process)
Table 12 - Green Premium baselines for the DAC pathways for 2020, indicative Green Premium against
PCC in italics.

The social cost of carbon (SCC) can be another reference point to the Green Premium
reductions. The social cost of carbon is the marginal cost of the impacts caused by emitting
one extra metric ton of greenhouse gas (CO2e) at any point in time (measuring the impacts on
the environment and human health). Although SCC has been used in appraising the impacts
of project (cost-benefit analysis) in multiple jurisdictions (e.g., US, UK, EU, Canada) the use
of it has been challenged by lawmakers. The reason is that the value of SCC is highly
uncertain, and it is sensitive to assumptions, such as future emissions, climate response, and
climate damages. The SCC is a progressively increasing value, its value depends on the year
when the emissions take place, the assumed value of future damages, and the stringency of
climate regulation. Here we used average values from US federal 15 and UK Department for
Environment, Food and Rural Affairs (Defra) values (Watkiss et al., 2006). Green Premium vs
the social cost of carbon shown in Table 13.
Levelized costs of capture
Technology HT-DAC ($/tCO2) LT-DAC ($/tCO2) SCC ($/tCO2)
Price ceiling 550 730 482
Average price 243 303 71
Green Premium
172 232 n.a.
against SCC
Table 13 - Green Premium baselines for the DAC pathways for 2020, indicative Green Premium against
SCC in italics.

Investees should report levelized-costs for their operations in line with the methods
described above. These values can be used for future reference in calculating Green
Premium reductions based on actual capture cost data, although further analysis and data
collection is recommended looking at both inter-annual variation of the Green Premium as

15
https://www.scientificamerican.com/article/cost-of-carbon-pollution-pegged-at-51-a-ton/
33
33
well as regional variation, which will be done as soon as data is available. In Figure 13 cost
reductions are shown with and without catalytic investment. The assumptions are as follow:

 Catalytic investment: $500,000,000


 Learning rate: 15%
 Demand elasticity: 1.57
 Effective investment proportion for learning and cost reduction: 82.4%

Figure 13 - Indicative cost reduction and Green Premium reduction due to additional investment ($500 M
during 2 years) for the HT-DAC pathway

34
34
Uncertainty and challenges
There is considerable uncertainty built in the cost model, particularly on what respects key
variables such as future prices of energy and materials (costs) as well as price of final
production (revenues), working hours and production volumes and cost of capital. Overall
uncertainty affecting these values will come from both the variability of prices with time and
geography and unpredictability of prices with time. In terms of financial analysis and risk, this
would typically be dealt through the construction of price scenarios. We recommend that cost
models should be built and tested to their overall sensitiveness to key parameters and
variables.
The analysis of existing literature and modelling exercises also revealed that cost uncertainty
is higher in some technological pathways than others – partially reflecting their different
technological maturities and uncertainties on input prices, geographical differences also play
a considerable role.
Furthermore, despite a certain ubiquity of learning curves to model technology cost reductions,
it is extremely difficult to find reliable references and data to model learning curves. This is an
area where future data that might be produced by investments could help reduce uncertainty
of the current modeling approach.

35
35
Catalyzed emissions reductions
As per the ECT Framework, to calculate the CatER use the formula:
2050 2050

𝐶𝑎𝑡𝐸𝑅𝑛 = ∑ 𝐶𝑎𝑡𝐸𝑅𝑦 = ∑ 𝐶𝐴𝐹𝑦 ∗ 𝐶𝑎𝑡𝑎𝑙𝑦𝑠𝑒𝑑 𝐴𝑐𝑡𝑖𝑣𝑖𝑡𝑦𝑦


𝑦 𝑦

with
𝐶𝑎𝑡𝑎𝑙𝑦𝑧𝑒𝑑 𝐴𝑐𝑡𝑖𝑣𝑖𝑡𝑦𝑦
= 𝐴𝑐𝑡𝑖𝑣𝑖𝑡𝑦 𝑣𝑜𝑙𝑢𝑚𝑒𝐶𝑎𝑡𝑎𝑙𝑦𝑠𝑒𝑑 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 𝑠𝑐𝑒𝑛𝑎𝑟𝑖𝑜
− 𝐴𝑐𝑡𝑖𝑣𝑖𝑡𝑦 𝑣𝑜𝑙𝑢𝑚𝑒𝐵𝑎𝑠𝑒𝑙𝑖𝑛𝑒 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 𝑠𝑐𝑒𝑛𝑎𝑟𝑖𝑜

Where:
𝐶𝑎𝑡𝐸𝑅𝑛 : Catalyzed Emissions Reduction during n years (n=30)
𝐶𝑎𝑡𝐸𝑅𝑦 : Catalyzed Emissions Reduction in year y

𝐶𝐴𝐹𝑦 : Carbon abatement factor in year y

The cumulative values of the ERP (or potential avoided emissions), including CatER according
to the above, can be seen in Table 15, Figure 14, and in Figure 15. The below results assume:

 Investment of $500,000,000 made in 2022


 Pathway learning rates
o HT-DAC: 15%
o LT-DAC: 15%
 Effective investment proportion for learning
o HT-DAC: 82.4%
o LT-DAC: 82.5%
 50% market share for each pathway in the available DAC market (see Table 14 for
activity volumes)
Refernce scenario Pathway activity Pathway activity
activity volumes volumes - reference volumes –
(MtCO2) scenario (50% of accelerated scenario
market) (50% of market)
(MtCO2) (MtCO2)
2021 0.238 0.119 0.119
2022 1.089 0.545 0.545
2023 2.575 1.288 3.486
2024 4.611 2.306 6.167
2025 7.128 3.564 7.963
2026 10.075 5.038 9.931
2027 13.415 6.707 12.060
2028 17.116 8.558 14.342
2029 21.155 10.577 16.767
2030 25.510 12.755 19.330
2031 30.164 15.082 22.023
2032 35.102 17.551 24.841

36
36
2033 40.309 20.155 27.780
2034 45.775 22.887 30.834
2035 51.487 25.744 33.999
2036 57.437 28.719 37.271
2037 63.615 31.808 40.648
2038 70.013 35.007 44.125
2039 76.624 38.312 47.699
2040 83.440 41.720 51.367
2041 90.455 45.227 55.127
2042 97.662 48.831 58.975
2043 105.056 52.528 62.910
2044 112.631 56.316 66.929
2045 120.383 60.191 71.029
2046 128.305 64.153 75.210
2047 136.395 68.197 79.467
2048 144.646 72.323 83.801
2049 153.055 76.527 88.208
2050 161.618 80.809 92.686
Table 14 - Activity volumes in the Reference and the Catalysed scenario (50% of market share as per
pathway activity volumes)

Catalyzed Emissions
Reduction Potential until 2050
(MMtCO2)
HT-DAC 53.98
LT-DAC 65.96
Table 15 - Emissions reduction potential of the two DAC pathways

Figure 14 - Annual emissions reduction potential for the High Temperature DAC (HT-DAC) pathway

37
37
Figure 15 - Annual emissions reduction potential for the Low Temperature DAC (LT-DAC) pathway

Uncertainty and challenges


The CatER figures have the same type of uncertainties that characterize ERP, namely the
uncertainty of Carbon Abatement Factors. Additionally, CatER are highly dependent of the
market uptake scenario, as well as the estimates of investment able to accelerate the
deployment of the reference scenario.
To deal with this uncertainty, the best option is to consider different available scenarios and
this challenge of scenario selection and consideration of different scenarios, should be
addressed in a next iteration of this document.
In addition to this question, the shape of the scenario (given the same activity level in 2050)
also significantly influences the ERP, as it increases or decreases the amount of displacement
(to the left, this is, earlier in time) of the reference scenario for a given fixed investment amount.

38
38
References
APS (2011). Direct Air Capture of CO2 with Chemicals: A Technology Assessment for the
APS Panel on Public Affairs. Online. Available at:
https://www.aps.org/policy/reports/assessments/upload/dac2011.pdf
Beuttler, C., Charles, L. and Wurzbacher, J. (2019). The Role of Direct Air Capture in
Mitigation of Anthropogenic Greenhouse Gas Emissions. Frontiers in Climate. 1.
Breyer, C., Fasihi, M. and Aghahosseini, A. (2020). Carbon dioxide direct air capture for
effective climate change mitigation based on renewable electricity: a new type of energy
system sector coupling. Mitig Adapt Strateg Glob Change. 25, pp.43–65.
Brinckerhoff, P. 2011. Accelerating the uptake of CCS: industrial use of captured carbon
dioxide Report.
Broehm, M. Strefler, J. and Bauer, N. (2015). Techno-Economic Review of Direct Air Capture
Systems for Large Scale Mitigation of Atmospheric CO2.
Broehm, M., Strefler, J. and Bauer, N. (2015). Techno-Economic Review of Direct Air Capture
Systems for Large Scale Mitigation of Atmospheric CO2. Online. Available at:
https://www.researchgate.net/publication/314428770_Techno-
Economic_Review_of_Direct_Air_Capture_Systems_for_Large_Scale_Mitigation_of_Atmos
pheric_CO2
Bui, M., et al. (2018). Carbon capture and storage (CCS): the way forward. Energy Environ.
Sci. 11, p.1062.
Cuéllar-Franca, R. and Azapagic, A. (2015). Carbon capture, storage and utilisation
technologies: A critical analysis and comparison of their life cycle environmental impacts.
Journal of CO2 Utilization. 9, pp.82-102.
Chatterjee, S. and Huang, K.W. (2020). Unrealistic energy and materials requirement for direct
air capture in deep mitigation pathways. Nat Commun. 11, p.3287.
Chen, C. andTavoni, M. (2013). Direct air capture of CO2 and climate stabilization: A model
based assessment. Climatic Change. 118, pp.59–72.

Daggash, H.A., Heuberger, C.F., MacDowell, N. (2019). The role and value of negative
emissions technologies in decarbonising the UK energy system. International Journal of
Greenhouse Gas Control. 81, pp.181-198.
Deutz, S. and Bardow, A. (2020). How (Carbon) Negative Is Direct Air Capture? Life Cycle
Assessment of an Industrial Temperature-Vacuum Swing Adsorption Process. ChemRxiv.
Online. Available at: https://doi.org/10.26434/chemrxiv.12833747.v2
Fajardy, M., Koberle, A., MacDowell, N. and Fantuzzi, A. (2019). BECCS deployment: a reality
check. Grantham Institute Briefing Paper No 28.
Fasihi et al. (2010). Direct Air Capture of CO2: A Key Technology for Ambitious Climate
Change Mitigation. 3(9), pp.2053-2057.

Fasihi, M., Efimova, O. and Breyer, C. (2019). Techno-economic assessment of CO2 direct
air capture plants. Journal of Cleaner Production. 224(1), pp.957-980.

39
39
Gambhir, A. and Tavoni, M. (2020). Direct Air Carbon Capture and Sequestration: How It
Works and How It Could Contribute to Climate-Change Mitigation. One Earth. 1(4), pp.405-
409.
Giesen, C., Meinrenken, C.J., Kleijn, R., Sprecher, B., Lackner, K.S., and Kramer., G.J.
(2017). A Life Cycle Assessment Case Study of Coal-Fired Electricity Generation with
Humidity Swing Direct Air Capture of CO2 versus MEA-Based Postcombustion Capture.
Environ. Sci. Technol. 51(2), pp.1024–1034.
Ha-Duong, M. and Keith, D.W. (2002). Climate strategy with CO2 capture from the air. Climatic
Change.
Hanna, R., Abdulla, A., Xu, Y. et al. (2021) Emergency deployment of direct air capture as a
response to the climate crisis. Nat Commun. 12, p.368
Harper, A.B., Powell, T., Cox, P.M. et al. (2018). Land-use emissions play a critical role in
land-based mitigation for Paris climate targets. Nat Commun. 9, p.2938.
Hepburn, C., Adlen, E., Beddington, J. et al. (2019). The technological and economic
prospects for CO2 utilization and removal. Nature. 575, pp.87–97.
House, K.Z., Baclig, A.C., Ranjan, M., van Nierop, E.A., Wilcox, J. and Herzog, H.J. Economic
and energetic analysis of capturing CO2 from ambient air. Proceedings of the National
Academy of Sciences. 108(51), pp.20428-20433.
ICEF (2018). ICEF Roadmap. Online. Available at: https://www.globalccsinstitute.com/wp-
content/uploads/2020/06/JF_ICEF_DAC_Roadmap-20181207-1.pdf
IEA (2020). Tracking Energy Integration. IEA: Paris. Online. Available at:
https://www.iea.org/reports/tracking-energy-integration-2020
IEA (2020). World Energy Model. IEA, Paris. Online. Available at:
https://www.iea.org/reports/world-energy-model
IEA (2021). Net Zero by 2050. IEA, Paris. Online. Available at: https://www.iea.org/reports/net-
zero-by-2050
Ishimoto, Y. et al. (2017) Putting costs of direct air capture in context. FCEA Working Paper
Series. Online. Available at: https://ssrn.com/abstract=2982422
Iyer G., Hultman N., Eom J., McJeon H., Patel P. And Clarke L. (2013). Diffusion of low-carbon
technologies and the feasibility of long-term climate targets. Technological Forecasting and
Social Change. ISSN 0040-1625. Online. Available at:
http://dx.doi.org/10.1016/j.techfore.2013.08.025
Izikowitz, D. (2021). Carbon Purchase Agreements, Dactories, and Supply-Chain Innovation:
What Will It Take to Scale-Up Modular Direct Air Capture Technology to a Gigatonne Scale,
PERSPECTIVE. Online. Available at: https://doi.org/10.3389/fclim.2021.636657
Jonge, M.M.J., Daemen, J., Loriaux, J.M., Steinmann, Z. and Huijbregts, M. (2019). Life cycle
carbon efficiency of Direct Air Capture systems with strong hydroxide sorbents. Elsevier
International Journal of Greenhouse Gas Control Volume. 80, pp.25-31.
Keith, D., Heidel, K. and Cherry, R. (2010). Capturing CO2 from the atmosphere: Rationale
and Process Design Considerations. Online. Available at:
https://keith.seas.harvard.edu/publications/capturing-co2-atmosphere-rationale-and-process-
design-considerations
40
40
Keith, D.W. et al. (2018), A process for capturing CO2 from the atmosphere, Joule, Vol. 2, No.
8, Cell Press, Cambridge, MA, pp. 1573–1594, doi: 10.1016/J.JOULE.2018.05.006.
Kelemen, P., Benson, S.M., Pilorgé, H., Psarras, P. and Wilcox, J. (2019). An Overview of the
Status and Challenges of CO2 Storage in Minerals and Geological Formations. Front. Clim.
Available at: https://doi.org/10.3389/fclim.2019.00009
Kiani. A., Jiang, K. and Feron, P. (2020). Techno-Economic Assessment for CO2 Capture
From Air Using a Conventional Liquid-Based Absorption Process. Front. Energy Res. 8(92),
Kintisch, E. (2014). MIT technology review: can sucking CO2 out of the atmosphere really
work? Online. Available at: http://globalthermostat.com/wp-content/uploads/2014/10/MIT-
Technology-Review-Global-Thermostat.pdf
Kirchofer, A. et al. (2012). Impact of alkalinity sources on the life-cycle energy efficiency of
mineral carbonation technologies. Energy Environ. Sci. 5, pp.8631–41.
Kulkarni, A.R. And Sholl, D.S. (2012). Analysis of Equilibrium-Based TSA Processes for Direct
Capture of CO2 from Air. Ind. Eng. Chem. Res. 51(25), pp.8631–8645.
Lackner, K.S. et al. (1995)., CARBON DIOXIDE DISPOSAL IN CARBONATE MINERAL
Energy. 20(11), pp.1153-1170.
Lackner, K.S., Ziock, H. and Grimes, P. (1999). Carbon Dioxide Extraction from Air: Is It an
Option? Proceedings of the 24th Annual Technical Conference on Coal Utilization & Fuel
Systems. pp.885–896.
Lackner, K.S. (2009). Capture of carbon dioxide from ambient air. Eur. Phys. J. Special Topics.
176, pp.93–106.
Liu, C.M., a Navjot, K., Sandhu, B., McCoy, S.T. and Bergerson, J.A. (2020). A life cycle
assessment of greenhouse gas emissions from direct air capture and Fischer– Tropsch fuel
production. Sustainable Energy Fuels. 4, p.3129.
Mazzotti, M., Baciocchi, R., Desmond, M. J. & Socolow, R. H. (2013). Direct air capture of
CO2 with chemicals: optimization of a two-loop hydroxide carbonate system using a
countercurrent air-liquid contactor. Clim. Change. 118, pp.119–135
McQueen, N. and Wilcox, J. (2020). Technoeconomic Assessment Tool for Direct Air Capture
(DAC_TEA_User_Guide.3.01.20)
Meckling, J. and Biber, E.A. (2021) Policy roadmap for negative emissions using direct air
capture. Nat Commun. 12, p.2051.

Müller, L.J., Kätelhön, A., Bachmann, M., Zimmermann, A., Sternberg, A. and Bardow, A.
(2020). A Guideline for Life Cycle Assessment of Carbon Capture and Utilization. Front.
Energy Res. Available at: https://doi.org/10.3389/fenrg.2020.00015
NASEM (2019). Negative Emissions Technologies and Reliable Sequestration: A Research
Agenda.
Neil, G. et al (2021). Confronting mitigation deterrence in low-carbon scenarios, Environ. Res.
Lett. 16, pp.64-99.
Nemet, G.F. and Brandt, A.R. (2012). Willingness to pay for a climate backstop: liquid fuel
producers and direct CO2 air capture. Working Paper Series La Follette School Working Paper
No. 2011-015

41
41
Nemet, G. (2019. How Solar Energy Became Cheap: A Model for Low-Carbon Innovation.
Routledge.
Pradhan, S., Shobe, W.M., Fuhrman, J., McJeon, H., Binsted, M., Doney, S.C. and Clarens,
A.F. (2021). Effects of Direct Air Capture Technology Availability on Stranded Assets and
Committed Emissions in the Power Sector. Front. Clim. 3, p.660787.
Pradhan S.P., Shobe, W., Fuhrman, J., McJeon, H. et al. (2021). Effects of Direct Air Capture
Technology Availability on Stranded Assets and Committed Emissions in the Power Sector.
Frontiers in Climate. 3.
Ranjan, M., & Herzog, H. J. (2011). Feasibility of air capture. Energy Procedia. 4, pp.2869–
2876.
Ram M., Bogdanov, D., Aghahosseini, A., Gulagi A., et al. (2018). Global Energy System
based on 100% Renewable Energy – Energy Transition in Europe Across Power, Heat,
Transport and Desalination Sectors. Study by LUT University and Energy Watch Group,
Lappeenranta, Berlin.
Realmonte G., Drouet, L., Gambhir, A. et al. (2019). An inter-model assessment of the role of
direct air capture in deep mitigation pathways. Nat Commun. 10, 3277.
Rosa, R.N. (2017). The Role of Synthetic Fuels for a Carbon Neutral Economy. 3(2), p.11.
Rosa, L., Sanchez, D.L., Realmonte, G., Baldocchi, D. and D’Odorico, P. (2021) The water
footprint of carbon capture and storage technologies. Renewable and Sustainable Energy
Reviews. 138, p.110511.
Rosental, M., Fröhlich, T. and Liebich, A. (2020). Life Cycle Assessment of Carbon Capture
and Utilization for the Production of Large Volume Organic Chemicals. Front. Clim. 2,
p.586199.
Sanz-Pérez, E.S., Murdock, C.R., Didas, S.A. and Jones., C.W. (2016). Direct Capture of CO2
from Ambient Air. Chem. Rev. 116(19), pp.11840–11876.
SAPEA, Science Advice for Policy by European Academies. (2018). Novel carbon capture
and utilisation technologies: research and climate aspects. Berlin: SAPEA.
Socolow, R. et al. (2011). Direct air Capture of CO2 with chemicals. a technology assessment
for the APS panel on public affairs. Tech. Rep.
Stephens, A. and Thieme, V. (2020). The Avoided Emissions Framework (AEF).
Terlouw, T. Bauer, C., Rosa, L. and Mazzotti, M. (2021). Life cycle assessment of carbon
dioxide removal technologies: a critical review. Energy Environ. Sci. 14, p.1701.
Terlouw, T., Treyer, K., Bauer, C. and Mazzotti, M. (2021). Life Cycle Assessment of Direct
Air Carbon Capture and Storage with Low-Carbon Energy Sources. ChemRxiv. Online.
Available at: https://doi.org/10.26434/chemrxiv.14346182.v1
The Royal Society and Royal Academy of Engineering (2018). Greenhouse gas removal.
Issued: DES5563_1 ISBN: 978-1-78252-349-9
Thonemann, N. and Pizzol, M. (2019). Consequential life cycle assessment of carbon capture
and utilization technologies within the chemical industry. Energy Environ. Sci. 12, p.2253.
Watkiss, P. et al (2006). The Social Costs of Carbon (SCC) Review – Methodological
Approaches for Using SCC Estimates in Policy Assessment. Final Report to Defra. With
42
42
contributions from David Anthoff, Tom Downing, Cameron Hepburn, Chris Hope, Alistair Hunt,
and Richard Tol.
Wilcox, J., Psarras, P.C. and Liguori, S. (2017). Assessment of reasonable opportunities for
direct air capture Department of Chemical and Biological Engineering. Colorado School of
Mines. Golden, CO: United States of America.
Wurzbacher, J.A., Gebald, C., Brunner, S. and Steinfeld, A. (2016). Heat and mass transfer
of temperature–vacuum swing desorption for CO2 capture from air. Chemical Engineering
Journal. 283, pp.1329-1338.

43
43

You might also like