You are on page 1of 16

Mitigation and Adaptation Strategies for Global Change

https://doi.org/10.1007/s11027-019-09850-z

ORIGINAL ARTICLE

Techno-economic assessment of alternative fuels


in second-generation carbon capture
and storage processes

Martin Haaf 1 1 1
& Peter Ohlemüller & Jochen Ströhle & Bernd Epple
1

Received: 31 August 2018 / Accepted: 20 February 2019/


# Springer Nature B.V. 2019

Abstract
Several technical methods are currently discussed to meet the objectives of the United Nations
Framework Convention on Climate Change 21st Conference of the Parties, Paris, France
(Paris Agreement) in terms of carbon dioxide (CO2) concentration in the Earth’s atmosphere.
In addition to efficiency improvements, reduction of energy consumption, and the utilization
of renewable energy sources, the application of carbon capture and storage (CCS) technologies
seems to be unavoidable. Whereas all these measures aim on the reduction of CO2 that is being
newly released, there is the approach to remove CO2 from the atmosphere that has already
been emitted. This can be achieved by the utilization of bioenergy in CCS processes. Within
this paper, the utilization of alternative fuels in two second-generation CCS processes is
assessed. In this regard, chemical looping combustion (CLC) and calcium looping (CaL) are
two promising technologies. Both processes have proven their feasibility already in semi-
industrial scale. The assessment includes three different types of fuel namely coal, biomass,
and solid recovered fuel (SRF). The analysis is twofold: first, a heat and mass balance
calculation reveals the specific CO2 emissions of each power system; second, a cost analysis
points out the feasibility from an economic point of view. The highest CO2 removal can be
achieved by a biomass-fired CLC unit (− 696 gCO2/kWhe). Furthermore, it was found that the
co-combustion of SRF even at moderate co-firing rates allows for noteworthy improved
economics of the CCS system. Therefore, the utilization of waste-derived fuels in the context
of CCS processes should be put more into focus in future research activities.

Keywords CaL . CLC . CCS . BECCS . Alternative fuels . Economics

* Martin Haaf
martin.haaf@est.tu-darmstadt.de

1
Institute for Energy Systems and Technology, Technische Universität Darmstadt, Otto-Berndt-Straße
2, 64287 Darmstadt, Germany
Mitigation and Adaptation Strategies for Global Change

1 Introduction

Man-made climate change has been globally accepted and addressed as an economic
and environmental threat for mankind. In order to keep the rise of the Earth’s
temperature in acceptable levels, the carbon dioxide (CO2) concentration in the
atmosphere is of major concern. Among other greenhouse gases (GHGs), CO2 repre-
sents the major contributor to anthropogenic climate change. Several technical
methods are currently discussed to meet the objectives set by the Paris Agreement
in terms of CO2 concentration in the atmosphere. There is the general willingness to
limit the further increase of the global temperature below 2 °C compared to prein-
dustrial levels (IPCC 2014). In addition to efficiency improvements in all steps of
energy conversion processes, reduction of energy consumption and the utilization of
renewable sources for heat and power supply, the application of carbon capture and
storage (CCS) technologies seems to be unavoidable. Considering the fact that several
global climate scenarios raise the need for an effective reduction of the CO2 that is
already present in the atmosphere during the second half of this century, novel
approaches are required (IEA 2011).
Processes such as direct CO2 air capture (San-Perez et al. 2016) or the removal of CO2
due to the capture and subsequent storage of biogenic CO2 (BECCS) are proposed (IEA
2011; Kemper 2015). Biomass is usually treated as carbon-neutral, since the carbon has
been taken from the atmosphere during the growth of the feedstock by means of
photosynthesis. There are several estimations published by different research groups,
showing the potential of global bioenergy. Numbers from the Intergovernmental Panel on
Climate Change Fifth Assessment Report indicate a significant role for biomass, provid-
ing 5–95 EJ/year in 2030, 10–245 EJ/year in 2050, and 105–325 EJ/year in 2100 (IPCC
2014). Other studies determine the bioenergy potential of 50–500 EJ/year (IPCC 2011)
or 160–270 EJ/year (Johannsson et al. 2012). For the year 2050, the potential for
negative CO2 emissions is approximately 10 GtCO2/year for biomass co-firing in pulver-
ized coal (PC) CCS units and for dedicated biomass firing in circulating fluidized bed
(CFB) combustion systems. Figure 1 depicts a simplified carbon balance for different
energy systems with (w) or without (w/o) CCS, utilizing fossil and biogenic carbon
(IEAGHG 2011).
Fossil carbon is marked in black, whereas biogenic carbon is indicated in gray. The
CO2 emissions to the atmosphere are marked as arrows directed upwards, and the
downwards directed arrows represent the CO2 that is captured and sequestrated. In
today’s state-of-the art energy systems, the main share of heat is supplied by fossil-
based carbon sources, such as coal, oil, and natural gas (IEA 2015). The fuel is burnt,
and the CO2 formed during combustion is released to the atmosphere with the exhaust

fossil carbon biogenic carbon

Energy System Energy System Energy System Energy System


w/o CCS w CCS w/o CCS w CCS

case I: +100 % case II: +10 % case III: +/- 0% case IV: - 90%

Fig. 1 Simplified carbon balance for energy systems burning fossil and biogenic carbon
Mitigation and Adaptation Strategies for Global Change

gas stream. Thus, 100% of the carbon is emitted as CO2 to the atmosphere (case I). Up to
approximately 90% of the CO2 can be captured and sequestrated, if a CCS unit is
employed (case II). Thereby, only the remaining 10% of the CO2 are released to the
atmosphere, which is due to chemical and technical limitations in the process of CO2
capture. However, the overall CO2 balance is still positive even if a CCS system is
applied. The cases III and IV represent an energy system, which is supplied by biogenic
carbon. Carbon neutrality is feasible, if biomass is burnt exclusively (see case III). In
case IV, a negative carbon balance is prevailing, due to the capture of CO2 that is formed
during the combustion of biomass. This case represents the framework of this study, with
the focus of different types of fuel (i.e., different biogenic carbon content) and two
different second-generation CCS technologies, namely chemical looping combustion
(CLC) and calcium looping (CaL). Both processes are ranked at a technology readiness
level (TRL) of 6, which means both have been demonstrated at relevant scale in an
industrial environment (Abanades et al. 2015). Putting this into perspective, post-
combustion CO2 capture by means of amine scrubbing is leveled on TRL 9 (Bui et al.
2018); hereby, the actual system was proven in an operational environment. The appli-
cation of inorganic membranes for post-combustion CO2 capture on large scale is leveled
on TRL 2–3 (Pera-Titus 2013). This means the technology concept has been formulated
and the experimental proof of concept of key components at small bench scale was
successfully carried out.
CLC is a novel process for oxy-fuel combustion. Figure 2 shows a basic process
scheme of this technology. Two interconnected fluidized bed (FB) reactors, the air reactor
(AR) and the fuel reactor (FR), replace the combustion chamber of a conventional power
plant. A solid material, the so-called oxygen carrier (OC), is used for in situ air separation
that takes place as follows: The OC is oxidized by atmospheric oxygen in the AR. The
oxidized OC is transferred to the FR to convert the carbon-containing fuel subsequently.
Thus, there is no direct mixing of fuel and air-nitrogen. The main products of CLC are
oxygen-depleted air at the outlet of the AR and a highly concentrated CO2 stream at the
outlet of the FR (Adánez et al. 2018; Wang et al. 2015). After condensation of water vapor
and gas treatment, the CO 2 stream is ready for further handling (e.g., storage or
utilization).
The main CLC reactions are summarized below. In most applications, a metal oxide
(MexOy) constitutes the basic material of the OC. The heat of reaction depends on the OC
and the type of fuel. The oxidation of the OC with air is exothermic (Eq. 1), while the
conversion of solid fuel with the OC is endothermic in many cases (Eq. 2). However, CLC

O2 depleted air CO2-product

MexOy
Air Reactor MexOy-1 Fuel Reactor Steam

Air Fuel
Fig. 2 Process scheme of chemical looping combustion
Mitigation and Adaptation Strategies for Global Change

provides the same amount of heat as the corresponding combustion in conventional air
atmosphere.
2 Mex Oy−1 þ O2 →2 Mex Oy ð1Þ

 q 
Cp Hq Or þ 2p þ −r Mex Oy
2
q  q  ð2Þ
→p CO2 þ H2 O þ 2p þ −r 2 Mex Oy−1
2 2

Different research groups investigate CLC in laboratory scale (Shen et al. 2009; Abad et al.
2012; Markström et al. 2013; Thon et al. 2014; Ohlemüller et al. 2016; Ge et al. 2016).
Additionally, results from pilot plants up to MWth scale have been reported (Ohlemüller
et al. 2017; Chamberland and Edberg 2015). There are different concepts of CLC with solid
fuels. The best known are in situ gasification CLC (iG-CLC) and chemical looping with
oxygen uncoupling (CLOU) (Adanez et al. 2012). In the case of iG-CLC, the fuel is gasified
with steam or CO2 inside the FR. The iG-CLC process is well known, and there are various
CLC units based on this concept. The main challenge of this process is the optimal balance
between conversion of solids in the fuel reactor, conversion of volatiles as well as gasifi-
cation products. One way to improve the conversion of solids and gases is to use an OC with
oxygen uncoupling properties. This leads to higher conversion of solids and gases. On the
other hand, the lifetime of the OC could be lower due to decreasing activity caused by
impurities, such as ashes or sintering effects. The main advantage of CLC compared to other
CCS processes is the low efficiency penalty that could offer low CO2 avoidance cost. The
additional costs and the efficiency of a CLC unit using coal as fuel compared to a reference
CFB boiler were estimated by Lyngfeld and Leckner (Lyngfelt and Leckner 2015). The
main cost factors are the CO2 compression unit (10 EUR/tCO2) and the oxygen polishing step
(6.5 EUR/tCO2). Furthermore, the additional cost for the boiler, OC, fluidization devices, and
cost reduction due to low combustion air excess accumulates to approximately 3.5 EUR/
tCO2. The total estimate for the cost of CO2 avoided is 20 EUR/tCO2, with a corresponding
efficiency penalty of 3.9%-points. In another study, CLC was identified as one of the
cheapest carbon capture technologies (IPCC 2005). The potential of low CO2 avoidance
cost was also confirmed by a thermodynamic and economic study (Ohlemüller et al. 2018).
Accordingly, the major uncertainties are related to the determination of the costs for
equipment and installation. The CO2 avoidance cost was calculated in the range of 13–55
EUR/tCO2. The corresponding net electric efficiency was determined to 41% including the
compression of the highly concentrated CO2 stream.
The CaL process has been frequently discussed as a mean for post-combustion CO2
capture from fossil-fired power plants. In addition to the power sector, its implementation
in the field of cement making processes seems to be promising (Ozcan et al. 2013; Jordal
et al. 2017). This is with regard to the similarities in the type of supplementary material.
Both processes need natural limestone. The CaL process is suitable for being retrofitted to
existing units and for being considered in the design of newly built systems. The CaL
process makes use of two interconnected FB reactors. It has been first proposed in 1999
(Shimizu et al. 1999). Since then, the CaL process reached TRL 6, due to successful
demonstration in a relevant environment. In Fig. 3, a basic scheme of the standard CaL
process is shown.
Mitigation and Adaptation Strategies for Global Change

A solid sorbent stream is forced to circulation between two coupled FB reactors. The CO2
contained in the flue gas (FG) of the upstream combustion or chemical process, is directed
through the carbonator for CO2 absorption by the reversible reaction of CO2 and lime (Eq. 3).
CaOðsÞ þ CO2ðgÞ ↔CaCO3ðsÞ þ ΔH ΔH 25°C ¼ 178:2 kJ=mol ð3Þ

The CO2-depleted FG stream is released to the environment after particle cleanup and heat
recovery. The loaded sorbent is transferred to the calciner where its temperature is
increased up to 900 °C, and the CO2 is released due to the calcination of the limestone.
The required heat is supplied by oxy-fuel combustion of supplementary fuel in the
calciner. The air separation unit (ASU) for the on-site oxygen supply poses the main
auxiliary power consumer in a CaL system (Ströhle et al. 2009). Due to the decrease of the
CO2 carrying capacity of the circulating sorbent stream during operation, a constant feed
of fresh limestone (make-up) is required to keep the CO2 uptake in the carbonator on an
appropriate level. The ideal carbonator operating temperature of approximately 650 °C is
justified by equilibrium and kinetic limitation of the carbonation reaction. Due to the oxy-
fuel regime within the calciner, the temperature of the loaded sorbent stream needs to be
raised close to 900 °C to ensure an almost complete regeneration of the sorbent at
relatively high CO2 partial pressure. However, the deactivation of the sorbent is forced
along with higher operating temperature in the calciner. Many efforts have been made to
demonstrate the feasibility of the process in relevant conditions close to industrial appli-
cation. The first successful continuous operation of a CaL unit, yielding to CO2 capture
efficiencies up to 90% were reported from a 30 kWth unit at the Spanish National Research
Council (Rodriguez et al. 2011). Experimental findings from a 200 kWth unit sustained the
stable operation of CaL system with CO2 capture rates up to more than 80% (Dieter et al.
2013). The readiness of the process is proven by long-term pilot tests carried out in MWth
scale (Chang et al. 2014; Ströhle et al. 2014; Hilz et al. 2017). More than 1200 h of smooth
operation and CO2 capture rates up to 95% have been reported (Helbig et al. 2017; Hilz
et al. 2018). In addition to experimental investigations, thermodynamic process simula-
tions have been developed as a useful tool for the calculation of the process performance in
full-scale environment. Hereby, the excess heat of the CaL process is recovered in a state-
of-the-art supercritical water steam cycle, delivering additional electric power to the grid.
Numerous technical and economical full-scale retrofit studies have been carried out. The
expected efficiency drop ranges from 5 to 8%-points. This number highly depends on the
performance of the upstream host plant and the case-specific process layout. However, the
specific CO2 emissions range from 70 up to 120 kgCO2/kWhe, which represents a CO2
emission reduction of approximately 80–90%. The LCOE for power generation have been

CO2-depleted flue gas CO2-product

CaCO3
Fuel
Carbonator CaO Calciner O2
Make-up

Flue gas Purge


Fig. 3 Process scheme of the calcium looping process
Mitigation and Adaptation Strategies for Global Change

calculated to 40–80 EUR/MWhe, whereas the CO2 avoidance cost is determined to 25–50
EUR/tCO2 (Abanades et al. 2007; Romeo et al. 2009; Yang et al. 2010; Schaupp 2014;
Cormos 2014; Rolfe et al. 2018; Hanak et al. 2018).
In comparison, a techno-economic assessment of an amine-based CO2 capture process
retrofitted to a 650 MWe power plant, reveals an efficiency penalty of 10.6%-points,
whereas the LCOE accumulates to 131 $/MWhe and CO2-avoidance cost of 86 $/tCO2
(Li et al. 2016). Other studies determine the CO2-avoidance costs for amine-based post-
combustion CO2 capture in the range of 62–95 $/tCO2 (DOE 2007; Manzolini et al. 2015;
Raksajati et al. 2013).
This work gives an assessment of the utilization of alternative fuels in CaL and CLC
processes. Therefore, technical and economical key parameters are calculated and discussed.
The overall goal is to provide an outlook of future atmospheric CO2-removal potential with
regard to economic-economic characteristics.

2 Methodology

2.1 Technical evaluation

Figure 4 depicts the framework of the CLC and CaL assessment cases. The first case
comprises a newly built CLC power unit (Fig. 4a). Air together with fuel is introduced to
the system, whereas an O2-depleted air stream is released to the environment and a concen-
trated CO2 stream is available for storage or utilization. In parallel, electrical power is supplied
to the grid. In the second case, an existing pulverized coal power plant (PC-PP) is retrofitted by
a CaL system for post-combustion CO2 capture (Fig. 4b). Within this assessment case, fuel is
fed to PC-PP and to the calciner of the CaL unit. Additionally, the CaL system requires a
constant stream of make-up for the solid circulation system. The CO2-depleted FG stream is
released from the carbonator, whereas a highly concentrated CO2 stream is available at the
outlet of the calciner. The two independent power cycles of the host plant and the CaL system
deliver electricity.
In this study, hard coal represents the fuel of the PC-PP. In contrast, the fuel streams fed to
FR and calciner represents a fuel blend consisting of two different types of fuel (hard coal +
biomass, hard coal + SRF). Table 1 gives the composition of the three types of fuel that are
considered in this study.
The composition and consequently the actual share of biogenic carbon and the LHV of
SRF highly vary between countries, regions, and feedstock of the processing units for

a O2
depleted air
Concentrated
CO2
b Power
Power
Concetrated
CO2

Fuel 1
Fuel 1
CLC System Power Fuel 1 FG
Fuel 2 PC-PP CaL System Fuel 2
Limestone

Air Air CO2 depleted


Purge
FG

Fig. 4 Framework of the two assessment cases: CLC system (a) and PC-PP + CaL system (b)
Mitigation and Adaptation Strategies for Global Change

SRF. In the present study, a biogenic carbon share of 40% for SRF is assumed, which is in
accordance with other studies (Hilber et al. 2007; Del Zotto et al. 2015). Due to the
relatively high contents of sodium, potassium, and chlorine (Cl) in biomass and SRF,
special attention needs to be paid to material selection and operation conditions of the
boiler. To sustain the risk of fouling, solid deposition, and high-temperature corrosion,
lower steam parameters and dedicated materials are usually applied in the case of alter-
native fuels in single boiler units (Iacovidou et al. 2018; Teixeria et al. 2012). In the cases
of CLC and CaL energy systems, there is the possibility to extract high-grade heat from
gas streams that are not in direct contact with the unwanted species introduced by the fuel.
Thus, there is the potential of higher thermal efficiencies compared to single boiler
systems in the case of firing an alternative fuel. However, this assessment aims on a
general overview, rather than a detailed evaluation of heat integration opportunities.
Table 2 summarizes the key boundary conditions for the three assessment cases, which
are derived from thermodynamic process assessments using coal as fuel.
The CaL heat distribution ratio, ycomb, gives the share of the heat introduced to the
calciner with respect to the total heat duty of the energy system (PC-PP + CaL). In the
carbonator, a CO2 absorption rate of 80% is considered, whereas all the CO2 released in
the calciner is inherently captured. The specific CO2 emissions for the CLC system,
rCO2,CLC, are calculated according to Eq. (4), where ṁfuel is the the mass flow of the fuel
and xc is the corresponding carbon content in the fuel. The total thermal power is given by
Q̇.

M CO2
E CLC ⋅ ṁf uel ⋅ ⋅ xc
MC
rCO2;CLC ¼ ð4Þ
Q̇ ⋅ ηnet;CLC

The specific CO2 emissions for the CaL system, rCO2,CaL, are determined similarly according to
Eq. (5).

M CO2 M CO2
E carb ⋅ ṁf uel ⋅ ⋅ xc E calc ⋅ ṁf uel ⋅ ⋅ xc
MC MC
rCO2;CaL ¼ ⋅ ð1−ycomb Þ þ ⋅ ycomb ð5Þ
Q̇ ⋅ ηnet;PC PP Q̇ ⋅ ηnet;CaL

Table 1 Composition (wt.%), LHV (MJ/kg), and biogenic carbon content (%) of the fuels

Species Hard coal Biomass SRF

C 69.6 51.3 42.4


H 4.3 6.0 3.8
N 1.0 0.3 0.5
S 0.5 0.0 0.2
O 13.8 37.8 17.1
Cl 0.01 0.2 0.8
H2O 6.00 4.2 16.0
Ash 4.80 0.4 19.3
LHV 26.9 19.4 16.9
Biogenic carbon 0 100 40
Mitigation and Adaptation Strategies for Global Change

Table 2 Technical boundary conditions for the three assessment cases (reference, CaL, CLC)

Symbol Parameter Unit Value

ηnet,Ref Ref: net electric efficiency reference (%) 45.6


rCO2,Ref Ref: specific CO2 emissions (kg/kWhe) 0.746
Ecarb CaL: CO2 absorption efficiency carbonator (%) 80
Ecalz CaL: CO2 capture efficiency calciner (%) 100
ycomb CaL: heat distribution ratio (−) 0.5
Δηnet,CaL CaL: efficiency penalty (%-points) 6.8
Δηnet,CLC CLC: efficiency penalty (%-points) 4.6
ECLC CLC: CO2 capture rate (%) 90

For a fuel blend, the contribution of each type of fuel is determined according to its ratio in the
fuel blend. The biogenic-based CO2 that is formed and captured is treated as negative value in
the above equations. Equation (6) gives the specific CO2 avoidance, rCO2,av, for the CaL and
CLC case, respectively.
rCO2;av;i ¼ rCO2;i −rCO2;PC PP ð6Þ

2.2 Economical evaluation

The model for the economic evaluation is based on an approach described as follows
(Rubin and Booras 2012). The specific costs for power generation, k, are determined
according to Eq. (7). In this equation, a states the specific investment costs, α states the
annuity factor, and c states the specific operating costs. The operational time of the power
unit is calculated by the total hours of a year, TN, and the utilization factor, nA. The specific
heat demand, w, is the reciprocal of the electric efficiency of the system. The specific fuel
price is expressed by pw. Variable operating costs are stated by b.

a⋅αþc 1
k¼ ⋅ þ w ⋅ pW þ b ð7Þ
TN nA
The annuity factor, α, is calculated in Eq. (8) with a given interest factor, q, the system live
time, n, and the interest rate, i.
qn ⋅ ðq−1Þ
α¼ with q ¼ 1 þ i ð8Þ
qn −1
The specific operating costs per year, c, are calculated by Eq. (9). Hereby, nMA states the staff
of the unit, kp states the personal costs, and kIH states the maintenance rate.
nM A ⋅ k p
c¼ þ a ⋅ k IH ð9Þ
Pel
The CO2 avoidance costs, Kv, are determined according to Eq. (10).
k−k PC PP
Kv ¼ ð10Þ
rCO2;av

The boundary conditions used for the economic evaluation are summarized in Table 3.
Similar to the composition of SRF also its price varies strongly. For this study, a negative
Mitigation and Adaptation Strategies for Global Change

price of − 40 EUR/tSRF is assumed, which leads to a revenue per mass of SRF burnt. This
number depends on the regulatory waste management framework, gate fees and availabil-
ity of suitable waste streams. However, for the UK, a study showed a revenue per mass of
SRF utilized in the range of approximately 30–105 EUR/tSRF (Garg et al. 2009). For coal, a
price of 55 EUR/tcoal is assumed, as noted during July 2018 (Coal commodity 2018). The
cost for biomass depends on the required supply chain such as production, transport,
pretreatment, and the type of biomass that is utilized. For the present study, a biomass
price of 70 EUR/tbio is assumed, which is accordance with comparable studies (Koornneef
et al. 2012). Since fuel prices are of great concern, a dedicated sensitivity analysis is
carried out in Sect. 3.
In addition to the abovementioned general boundary conditions, specific economic bound-
ary conditions are introduced for the CLC and CaL assessment cases, respectively (see
Table 4). A sensitivity analysis shows the effect of the major economic boundary conditions
for this study (Sect. 3).

3 Results and discussion

This section includes the technical results of the assessment cases with the focus on the
total CO2 reduction potential and the specific CO2 emissions. Thereafter, the levelized cost
of electricity (LCOE) and the CO2 avoidance costs are presented and discussed. Further-
more, a sensitivity study is carried out to address the influence of the main boundary
conditions.

3.1 Technical assessment

The specific CO2 emissions for co-firing rates of biomass or SRF in the range of 0–100%
are shown in Fig. 5. The results are presented for the CaL and CLC cases, respectively.
The specific CO2 emission of the reference case (PC-PP) is marked with the black dot in
the figure. The case of 100% coal firing is equal to the case of 0% co-firing of secondary
fuel. The state of CO2 neutrality is marked with the dashed black line. For biomass (Bio)
and SRF, the co-firing rate is raised stepwise by 10%. For 100% coal-firing, the CLC
system emits 0.083 kg CO2/kWhe. This number is slightly higher in the CaL case
(0.086 kgCO2/kWhe) due to the lower net electric efficiency of a CaL unit. It comes
apparent that the influence of the co-firing rate in the CLC case is higher compared to the

Table 3 General boundary conditions for the economical evaluation (Junk 2017)

Symbol Parameter Unit Value

i Interest rate (%) 8


kp Yearly staff costs (EUR) 60,000
nMA Staff (−) 60
nA Yearly utilization (%) 85
n System lifetime (a) 25
pbiomass Price of biomass (EUR/t) 70
pSRF Price of SRF (EUR/t) − 40
pcoal Price of coal (EUR/t) 55
Mitigation and Adaptation Strategies for Global Change

Table 4 CLC and CaL specific economic boundary conditions (Ohlemüller et al. 2018; Rolfe et al. 2018)

Symbol Parameter Unit CLC CaL PC PP

a Specific investment costs (EUR/kWe) 2049 1778 1307


kIH Yearly maintenance rate (%) 1.5 2.0 1.5
b Other variable costs (ct/kWhe) 0.4 0.6 0.2

CaL case. This is due to the fact that the hard coal fueled PC-PP provides 50% of the
total thermal input throughout the whole analysis. Once the specific CO2 emission reach
a negative value, CO2 is removed from the atmosphere, while electric power is produced.
For the CLC case, co-firing rates of approximately 10 and 22% are sufficient to ensure
negative CO2 emissions for biomass and SRF, respectively. In the CaL case, co-firing
rates of approximately 20% for biomass firing and 40% for SRF firing are required for
the removal of atmospheric CO2. On a yearly basis, a full-scale biomass fired CLC unit
removes approximately 2.12 GtCO2, while the utilization of SRF in a CLC unit accounts
for a removal of 0.62 GtCO2. For the retrofit of an existing PC PP with the CaL process,
the yearly CO2 removal accumulates to approximately 1.07 GtCO2 and 0.20 GtCO2 in the
case of firing biomass or SRF, respectively.

3.2 Economic assessment

The higher costs of electricity are a major obstacle for a large-scale implementation of
CCS technologies. The CCS systems that are considered in this study are highly efficient
and less cost intense compared to standard CCS technologies. However, there is an

0.8

PC PP CaL SRF
0.6
CaL Bio CLC SRF
CO2-emi ssi on [kg CO2/ kWhe]

0.4 CLC Bio

0.2

-0.2

-0.4

-0.6

-0.8
0 20 40 60 80 100
share of secondary fuel [%]
Fig. 5 CO2 emissions of the CaL and CLC system utilizing SRF and biomass as secondary fuel
Mitigation and Adaptation Strategies for Global Change

increase in the capital as well as in the operating costs compared to a state-of-the-art


power plant without CCS. Competiveness of CCS applications might be achieved, once
the costs for CO2 avoidance pass the price for CO2 allowance. With regard to the
utilization of alternative fuels in CaL and CLC processes, there is a lack of data for
investment and maintenance costs in literature. Thus, there are three different economic
scenarios introduced for the following evaluation. The economic scenario Bbase^ refers
to boundary conditions based on coal-fired units, whereas an increase of 20 and 40% of
investment and maintenance costs are assumed for the corresponding share of co-firing,
respectively. Figure 6 shows the LCOE of the CaL case (Fig. 6a) and for the CLC case
(Fig. 6b) depending on the share of secondary fuel and the economic scenario. For the
sake of comparison, the LCOE of a PC PP are indicated as a dashed line additionally.
The CLC case allows for lower LCOE due to lower investment costs and a lower sorbent
demand. However, the differences between CaL and CLC cases for 100% hard coal
firing are negligible (56.97 EUR/MWhe vs 52.90 EUR/MWhe). The LCOE increases
along with a higher share of biomass co-firing for the CaL and CLC cases, respectively.
If SRF is utilized instead, there is a contrary trend towards lower LCOE for both cases.
For the CaL case, LCOE similar to those of power generation without CCS are feasible if
almost 80% of the total thermal input to the calciner is supplied by SRF. For the CLC
case, approximately 50% of SRF co-firing is required to reach the same LCOE of power
generation without CCS. With regard to the large-scale implementation of CCS systems,
this is especially interesting, since the drawback of high costs can be cushioned by the
utilization of SRF even at low co-firing rates. However, research is required to further
clarify the effects of SRF firing in CaL and CLC systems, both on a technical and
economical point of view.
Figure 7 shows the CO2 avoidance cost for the CaL case (Fig. 7a) and for the CLC
case (Fig. 7b). Similar to the previous section, the analysis is carried out depending on
the economic assessment scenario. The economic scenario Bbase^ refers to boundary

90 90
CaL SRF CaL Bio a CLC SRF CLC Bio
b
80 80

70 70
LCOE [EUR/ MWh e]

LCOE [EUR/ MWh e]

60 60
"40 %" "20 %" "base"
50 50 "40 %" "20 %" "base"

40 40
PC-PP PC-PP
30 30

20 20

10 10
0 20 40 60 80 100 0 20 40 60 80 100
share of secondary fuel [%] share of secondary fuel [%]

Fig. 6 LCOE of the CaL case (a) and CLC case (b) depending on the share of secondary fuel and economic
scenario. The economic scenario Bbase^ refers to boundary conditions of a coal-fired unit, whereas B20%^ and
B40%^ represent an increase in investment and maintenance cost by 20 or 40% for the corresponding share of
secondary fuel, respectively
Mitigation and Adaptation Strategies for Global Change

40 40
a b
30 30
CO2 avoi dance cos t [EUR/ t CO2]

CO2 avoi dance cos t [EUR/ t CO2]


20 20
"20 %" "40 %"
"base"
"20 %"
10 10 "40 %"
"base"
0 0

-10 -10

-20 -20
CaL SRF CaL Bio CLC SRF CLC Bio
-30 -30
0 20 40 60 80 100 0 20 40 60 80 100
share of secondary fuel [%] share of secondary fuel [%]

Fig. 7 CO2 avoidance costs for the CaL case (a) and CLC case (b) depending on the share of secondary fuel and
economic scenario. The economic scenario Bbase^ refers to the boundary conditions of a coal-fired unit, whereas
B20%^ and B40%^ represent an increase in investment and maintenance costs for the corresponding share of co-
firing, respectively

conditions based on coal-fired units, whereas an increase of 20 and 40% of investment


and maintenance costs are assumed for the corresponding share of co-firing, respectively.
The dashed black line indicates the CO2 European Emission Allowances (European CO2
Emission Allowances 2018) of approximately 20 EUR/tCO2.
For the CaL and CLC case, the higher fuel costs for biomass in contrast to coal are nearly
balanced by the additional amount of biogenic CO2, which is captured by the CCS unit. If SRF
is utilized instead of biomass, the costs for CO2 avoidance decrease down to negative values.
For the CaL case and the 20% economic scenario, the CO2 avoidance cost are approximately
− 5 EUR/tCO2. This number is further reduced down to − 20 EUR/tCO2 by the CLC case under
the same economic scenario.
For the utilization of SRF, the CO2 avoidance costs are highly influenced by fuel
composition and fuel price. In order to show the influence of the fuel price on the
LCOE and the CO2-avoidance costs, a sensitivity study is carried out. The fuel price is
altered by ± 30% based on the values listed in Table 2. Figure 8 depicts the results of
the sensitivity analysis. The CaL case is shown in graph (Fig. 8a), whereas graph (Fig.
8b) shows the results for the CLC case. The symbols give the type of fuel as well as
the fuel price that is considered. Towards the bottom line of the figure, the SRF co-
firing rate, and towards the right of each graph, the co-firing rate of biomass increases,
respectively.
There are clear trends for the LCOE and for the CO2 avoidance costs similar for both
CCS technologies. If biomass is utilized as secondary fuel, the LCOE increases along with
constant or slightly decreased CO2 avoidance costs. For the utilization of SRF as second-
ary fuel, the LCOE and the CO2 avoidance cost decreases. The trends are more robust for
the CLC case, since within the CaL case still 50% of the thermal duty is supplied by hard
coal. For the 100% biomass CLC case, the LCOE varies by approximately ± 10 EUR/
MWhe if the fuel price is altered by ± 30%. However, the co-firing of biomass in both
cases leads to lower CO2-avoidance costs, even if the price rises by 30%.
Mitigation and Adaptation Strategies for Global Change

40 40
a b
coal
30 100 % 30
100 % coal
ss
ma
CO2-avoi dance cost [EUR/ t CO2]

CO2-avoi dance cost [EUR/ t CO2]


20 bio 20
% ass
100 % biom
10 10 100

0 0
-30% SRF
-10 % SRF -10 -30% Bio
100 base SRF
-20 -30% SRF -30% Bio -20 base Bio
base SRF base Bio +30% SRF
-30 -30
+30% SRF +30% Bio SRF +30% Bio
100 %
-40 -40
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
LCOE [EUR/MW h e ] LCOE [EUR/MW h e ]

Fig. 8 Sensitivity analysis of the fuel price, CaL case (a) and CLC case (b)

4 Conclusion

This study presents an assessment of the potential of CLC- and CaL-based power
systems to reduce the CO2 emissions from the power sector by the utilization of
alternative fuels. Due to biogenic fractions of SRF and biomass, CO2 could be even
removed from the atmosphere. The highest potential for negative CO2 emissions (−
696 gCO2/kWhe) shows a biomass fueled CLC unit due to the expense of relatively
high LCOE. Cost savings in CCS application can be achieved by the utilization of
SRF. The utilization of this waste derived fuel could lead to LCOE in the range of
conventional coal based power generation. This is the case for 40% co-firing of SRF
in a CLC unit or 80% co-firing of SRF in the CaL case. Negative CO2 emissions are
feasible already at relatively low co-firing rates of alternative fuels, such as 40% SRF
co-firing in a CaL unit or 10% SRF co-firing for the case of a biomass-fired CLC
unit. For a CaL retrofit of an existing fossil-fired PC-PP, the CO2 balance can be
noteworthy reduced down to negative values. This is of particular interest since
nowadays there are still new coal-based power plants under construction. This anal-
ysis clearly highlights the ability of second-generation based power systems to
contribute effectively to a net removal of CO2 from the atmosphere. With regard to
economics, CLC- and CaL-based processes are still cost intense; however, once the
price for CO2 emissions allowance increases, the application becomes more expedi-
ently. As a main finding of this study, the utilization of SRF as supplementary fuel
could potentially compensate for higher investment costs. However, in order to draw
clear conclusion, further research is required in order to confirm the utilization of
waste derived fuels in second-generation CCS systems from a technical and econom-
ical point of view.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.
Mitigation and Adaptation Strategies for Global Change

References

Abad A, Adánez-Rubio I, Gayán P, García-Labiano F, Diego LF, Adánez J (2012) Demonstration of chemical-
looping with oxygen uncoupling (CLOU) process in a 1.5 kWth continuously operating unit using a Cu-
based oxygen-carrier. Int J Greenhouse Gas Control 6:189–200
Abanades JC, Grasa G, Alonso M, Rodriguez N, Anthony EJ, Romeo LM (2007) Cost structure of a post-
combustion CO2 capture system using CaO. Environ Sci Technol 41:5523–5527
Abanades JC, Arias B, Lyngfeld A, Mattisson T, Wiley DE, Li H, Ho MT, Mangano E, Brandani S (2015)
Emerging CO2 capture systems. Int J Greenhouse Gas Control 40:126–166
Adanez J, Abad A, Garcia-Labiano F, Gayan P, de Diego LF (2012) Progress in chemical-looping combustion
and reforming technologies. Prog Energy Combust Sci 38:215–282
Adánez J, Abad A, Mendiara P, Gayan P, Diego LF, Garcia-Labiano F (2018) Chemical looping combustion of
solid fuels. Prog Energy Combust Sci 65:6–66
Bui M, Adjiman CS, Bardow A, Anthony EJ, Boston A, Brown S, Fennell PS, Fuss S, Galindo A, Hackett LA,
Hallett JP, Herzog HJ, Jackson G, Kemper J, Krevor S, Maitland GC, Matuszewski M, Metcalfe IS, Petit C,
Puxty G, Reimer J, Reiner DM, Rubin ES, Scott SA, Shah N, Smit B, Trusler JPM, Webley P, Wilcox J,
Dowell NM (2018) Carbon capture and storage (CCS): the way forward. Energy Environ Sci 11:1062–1176
Chamberland, Andrus R H, Edberg C (2015) Alstom’s Chemical looping combustion technology with CO2
capture for new and existing coal-fired power plants. U.S. DOE/NETL CO2 Capture Technology Meeting
Chang MH, Chen WC, Huang CM, Lui WH, Chou YC, Chang WC, Chen W, Cheng JY, Huang KW, Hsu HW
(2014) Design and experimental testing of a 1.9 MWth calcium looping pilot plant. Energy Procedia 63:
2100–2108
Coal commodity stock exchange (2018) https://www.finanzen.net/rohstoffe/kohlepreis/euro. Accessed 27 August
2018
Cormos CC (2014) Economic evaluation of coal-based combustion and gasification power plants with post-
combustion CO2 capture using calcium looping cycle. Energy 78:665–673
Del Zotto L, Tallini A, Di Simone G, Molinari G, Cedola L (2015) Energy enhancement of solid recovered fuel
within systems of conventional thermal power generation. Energy Procedia 81:319–338
Dieter H, Hawthorne C, Zieba M, Scheffknecht G (2013) Progress in calcium looping post combustion CO2
capture: successful pilot scale demonstration. Energy Procedia 37:46–56
DOE/NETL (2007) Carbon dioxide capture from existing coal-fired power station. U. S Department of Energy,
National Energy Technology Laboratory
European CO2 Emission Allowances (2018) https://www.finanzen.net/rohstoffe/co2-emissionsrechte. Accessed
27 August 2018
Garg A, Smith R, Hill D, Longhurst PJ, Pollard SJT, Simms NJ (2009) An integrated appraisal of energy
recovery options in the United Kingdom using solid recovered fuel derived from municipal solid waste.
Waste Manag 29:2289–2297
Ge H, Guo W, Shen L, Song T, Xiao J (2016) Biomass gasification using chemical looping in a 25 kWth reactor
with natural hematite as oxygen carrier. Chem Eng J 286:174–183
Hanak DP, Eranz M, Nabavi SA, Jeremias M, Romeo LM, Manovic V (2018) Technical and economic feasibility
evaluation of calcium looping with no CO2 recirculation. Chem Eng J 335:763–773
Helbig M, Hilz J, Haaf M, Daikeler A, Ströhle J, Epple B (2017) Long-term carbonate looping testing in a 1
MWth pilot plant with hard coal and lignite. Energy Procedia 114:179–190
Hilber T, Maier J, Scheffknecht G, Agraniotis M, Grammelis P, Kakaras E, Glorius T, Becker U, Derichs W,
Schiffer HP, De Jong M, Torri L (2007) Advantages and possibilities of solid recovered fuel cocombustion in
the European energy sector. J Air Waste Manage Assoc 57(10):1178–1189
Hilz J, Helbig M, Haaf M, Daikeler A, Ströhle J, Epple B (2017) Long-term pilot testing of the carbonate looping
process in 1 MWth scale. Fuel 210:892–899
Hilz J, Helbig M, Haaf M, Daikeler A, Ströhle J, Epple B (2018) Investigation of the fuel influence on the
carbonate looping process in 1 MWth scale. Fuel Process Technol 169:170–177
Iacovidou E, Hahladakis J, Deans I, Velis C, Purnell (2018) Technical properties of biomass and solid recovered
fuel (SRF) co-fired with coal: impact on multi-dimensional resource recovery value. Waste Manag 73:535–
545
IEA (2011) Combining bioenergy with CCS. Reporting and accounting for negative emissions under UNFCCC
and Kyoto protocol. OECD/IEA, Paris, France
IEA (2015) CO2 emissions from fuel combustion. Paris, France
IEAGHG (2011) Potential for biomass and carbon dioxide capture and storage. Report 2011/06. IEA Greenhouse
R&D Programme, Cheltenham, UK
Mitigation and Adaptation Strategies for Global Change

IPCC (2005) Special report on carbon dioxide capture and storage. Cambridge University Press, Cambridge,
United Kingdom and New York, USA
IPCC (2011) IPCC special report on renewable energy sources and climate change mitigation. Prepared by
Working Group III of the Intergovernmental Panel on Climate Change. Cambridge University Press,
Cambridge, United Kingdom and New York, USA
IPCC (2014) Climate change 2014: mitigation of climate change. Working Group III Contribution to the Fifth
Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press,
Cambridge, United Kingdom
Johannsson T, Nakicenovic N, Patwardhan A, Gomez-Echeverri L, Turkenburg W (2012) Summary for
policymakers. Global energy assessment—towards a sustainable future (eds Johansson TB, Nakicenovic
N, Patwardhan A, Gomez-Echeverri L), pp. 3-30. Cambridge University press, International Institute for
Applied System Analysis, Laxenburg, Austria, Cambridge, UK and New York, NY
Jordal K, Voldsund M, Storset S, Fleiger K, Ruppert J, Spörl R, Hornberger M, Cinti G (2017) CEMCAP—
making CO2 capture retrofittable to cement plants. Energy Procedia 114:6175–6180
Junk M (2017) Technical and economical assessment of various carbonate looping process configurations.
Dissertation, Technische Universität Darmstadt. Cuvillier Verlag, ISBN: 9783736994485
Kemper J (2015) Biomass and carbon dioxide capture and storage: a review. Int J Greenhouse Gas Control 40:
401–430
Koornneef JV, Breevoort P, Hamelinck C, Hendriks C, Hoogwijk M, Koop K, Koper M, Dixon T, Camps A
(2012) Global potential for biomass and carbon capture dioxide capture, transport and storage up to 2050. Int
J Greenhouse Gas Control 11:117–132
Li K, Leigh W, Feron P, Yu H, Tade M (2016) Systematic study of aqueous monoethanolamine (MEA)-based
CO2 capture process: techno-economic assessment of the MEA process and its improvements. Appl Energy
165:648–659
Lyngfelt A, Leckner B (2015) A 1000 MWth boiler for chemical-looping combustion of solid fuels—discussion
of design and costs. Appl Energy 157:475–487
Manzolini G, Fernandez ES, Rezvani S, Macchi E, Goetheer ELV, Vlugt TJH (2015) Economic assessment of
novel amine based CO2 capture technologies integrated in power plants based on European Benchmarking
Task Force methodology. Appl Energy 138:546–558
Markström P, Linderholm C, Lyngfelt A (2013) Chemical-looping combustion of solid fuels—design and
operation of a 100 kW unit with bituminous coal. Int J Greenhouse Gas Control 15:150–162
Ohlemüller P, Busch JP, Reitz M, Ströhle J, Epple B (2016) Chemical-looping combustion of hard coal:
autothermal operation of a 1 MWth pilot plant. J Energy Resourc Technol 138:042203-042203-7
Ohlemüller P, Ströhle J, Epple B (2017) Chemical looping combustion of hard coal and torrefied biomass in a 1
MWth pilot plant. Int J Greenhouse Gas Control 65:149–159
Ohlemüller P, Olausson M, John M, Alobaid F, Ströhle J, Epple B (2018) Thermodynamic and economic
evaluation of a full-scale chemical looping plant. 5th International Conference on Chemical Looping, Utah,
USA
Ozcan DC, Hyungwoong A, Brandani S (2013) Process integration of a Ca-looping carbon capture process in a
cement plant. Int J Greenhouse Gas Control 19:530–540
Pera-Titus M (2013) Porous inorganic membranes for CO2 capture: present and prospects. Chem Rev 114:1413–
1492
Raksajati A, Ho MT, Wiley DE (2013) Reducing the cost of CO2 capture from flue gases using aqueous chemical
absorption. Ind Eng Chem Res 52:16887–16901
Rodriguez N, Alonso M, Abanades JC (2011) Experimental investigations of a circulating fluidized-bed reactor
to capture CO2 with CaO. AICHE J 57:1356–1366
Rolfe A, Huang Y, Haaf M, Rezvani S, McIlveen-Wright D, Hewitt NJ (2018) Integration of the calcium
carbonate looping process into an existing pulverized coal-fired power plant for CO2 capture: techno-
economic and environmental evaluation. Appl Energy 222:169–179
Romeo LM, Lara Y, Lisbona P, Martìnez A (2009) Economical assessment of competitive enhanced limestones
for CO2 capture cycles in power plants. Fuel Process Technol 90:803–811
Rubin ES, Booras G (2012) Toward a common method of cost estimation for CO2 capture and storage.
Proceedings of CCS Cost Workshop, Palo Alto, USA
San-Perez AS, Murdock RC, Didas AS, Jones WC (2016) Direct capture of CO2 from ambient air. Chem Rev
116:11840–11876
Schaupp D (2014) Economic analysis of the calcium looping process. CAL-MOD Workshop. https://www.pre-
sustainability.com/download/DatabaseManualMethods.pdf. Accessed 27 August 2018
Shen L, Wu J, Xiao J, Song Q, Xiao R (2009) Chemical-looping combustion of biomass in a 10 kWth reactor
with Iron oxide as an oxygen carrier. Energy Fuel 23:2498–2505
Mitigation and Adaptation Strategies for Global Change

Shimizu T, Hirama T, Hosoda H, Kitano K, Inagaki M, Tejima K (1999) A twin fluid-bed reactor for removal of
CO2 from combustion processes. Institution of Chemical Engineers 77:62–68
Ströhle J, Lasheras A, Galloy A, Eppe B (2009) Simulation of the carbonate looping process for post-combustion
CO2 capture from a coal-fired power plant. Chem Eng Technol 32:435–442
Ströhle J, Junk M, Kremer J, Galloy A, Epple B (2014) Carbonate looping experiments in a 1 MWth pilot plant
and model validation. Fuel 127:13–22
Teixeria P, Lopes H, Gulyurtlu I, Lapa N, Abelha P (2012) Evaluation of slagging and fouling tendency during
biomass co-firing with coal in a fluidized bed. Biomass Bioenergy 39:192–203
Thon A, Kramp M, Hartge EU, Heinrich S, Werther J (2014) Operational experience with a system of coupled
fluidized beds for chemical looping combustion of solid fuels using ilmenite as oxygen carrier. Appl Energy
118:309–317
Wang P, Means N, Shekhawat D, Berry D, Massoudi M (2015) Chemical-looping combustion and gasification of
coals and oxygen carrier development: a brief review. Energies 8:10605–10635
Yang Y, Rongrong Z, Liqiang D, Kavosh M, Patchigolla K, Oakey J (2010) Integration and evaluation of a power
plant with a CaO based CO2 capture system. Int J Greenhouse Gas Control 4:603–612

You might also like