You are on page 1of 9

Fuel Processing Technology 210 (2020) 106557

Contents lists available at ScienceDirect

Fuel Processing Technology


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

Research article

Experimental investigation of the carbonation reactor in a tail-end Calcium T


Looping configuration for CO2 capture from cement plants
Matthias Hornberger , Joseba Moreno, Max Schmid, Günter Scheffknecht

Institute of Combustion and Power Plant Technology (IFK), University of Stuttgart, Pfaffenwaldring 23, 70569 Stuttgart, Germany

ARTICLE INFO ABSTRACT

Keywords: The calcium looping process is a high temperature post-combustion CO2 capture technology that is expected to
CO2 capture be especially suitable for cement plants. Both processes rely on CaCO3 as common feedstock, enabling the
CCS reutilisation of purged calcium looping sorbent in the cement clinker manufacturing process. Thus, setting new
Calcium looping boundary conditions for the calcium looping process. So far, calcium looping CO2 capture has been demonstrated
Cement industry
at semi industrial scale for fossil fuelled power plants but not yet for cement plants. In this work, results obtained
Carbonator
from University of Stuttgart's 200 kW calcium looping pilot plant investigating the so called tail-end calcium
looping cement plant integration are presented focusing on the carbonation reactor. Different integration levels
between the cement plant and the calcium looping process were assessed investigating make-up ratios up to
0.9 mol mol−1 and CO2 concentrations up to 0.33 m3 m−3. Operation at high sorbent make-up rates or high
integration levels enhanced the sorbent's CO2 carrying capacity whereas for low integration levels the sorbent
activity converged towards its residual CO2 carrying capacity. For high sorbent make-up flows, CO2 capture was
limited by the carbonation equilibrium yielding CO2 capture rates as high as 98% in the carbonator at carbo­
nation temperatures around 600 ∘C, whereas for low integration CO2 capture was limited by the active amount of
CaO being fed to the carbonator. The active space time carbonator model was applied with satisfactory agree­
ment to the conducted experiments indicating the model's validity for calcium looping CO2 capture from cement
plant. The obtained results emphasise the suitability of calcium looping CO2 capture for the decarbonisation of
the cement sector.

1. Introduction are attributed to improved energy efficiency, switching to alternative


fuels and reducing the clinker to cement ratio. A well suited CCS
The cement sector is one of the largest industrial CO2 emission technology for the cement sector is the calcium looping CO2 capture
sources, emitting annually approx. 2.2 Gt which represents around 7% process. Originally proposed in 1999 by Shimizu et al. [3] it was in­
of global anthropogenic CO2 emissions [1]. In contrast to other energy tensively developed, investigated and demonstrated at scales up to
intensive industries such as power generation or steel production, the 1.7 MW focusing on the decarbonisation of fossil fired power plants
cement sector cannot be decarbonised by switching to alternative en­ over the recent decades [4–6]. By means of cyclic carbonation and
ergy sources such as ‘biogenic fuels', ‘renewable electricity’ or ‘green calcination of a calcium containing sorbent CO2 can be separated from a
hydrogen’, since a major share of a cement plant's CO2 emissions are flue gas and released into a concentrated CO2 stream. In a first step, CO2
inherent to its clinker production [1,2]. Around two thirds of a cement from the flue gas reacts exothermically with CaO at temperatures ran­
plant's CO2 emission originate from the calcination of carbonates pre­ ging from 600 ∘C to 700 ∘C in the so called carbonator (Eq. (1)). The
sent in the cement raw meal, whereas the remaining third is attributed formed CaCO3 is then transferred to a second reactor, the calciner, to be
to the actual clinker burning process [2]. Consequently, post-combus­ calcined at temperatures above 900 ∘C at oxy-fuel conditions, hence,
tion or oxy-fuel CCS technologies need to be employed to decarbonise releasing CO2 and providing CaO for repetitive CO2 absorption. The
the cement sector [1,2]. The IEA [1] expects CCS technologies to con­ energy required to drive the endothermic calcination reaction is usually
tribute to cumulative CO2 savings of 48% by 2050, while the other 52% provided by oxy-fuel combustion of an auxiliary fuel.

Abbreviations: BFB, Bubbling fluidised bed; CaL, Calcium looping; CFB, Circulating fluidised bed; CSTR, Continuous stirred-tank reactor; waf, water and ash free; wf,
water free; ad, air dried; SPECCA, specific primary energy cost of CO2 avoided

Corresponding author.
E-mail address: matthias.hornberger@ifk.uni-stuttgart.de (M. Hornberger).

https://doi.org/10.1016/j.fuproc.2020.106557
Received 2 April 2020; Received in revised form 23 July 2020; Accepted 10 August 2020
Available online 25 August 2020
0378-3820/ © 2020 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
M. Hornberger, et al. Fuel Processing Technology 210 (2020) 106557

Nomenclature (mol s−1)


NCa, Clinker Molar flow of calcium required to produce clinker (mol
EA Activation energy (kJ mol−1) s−1)
ECO2, Carb, eq Equilibrium CO2 capture efficiency (mol mol−1) NCaCO3,0 Molar flow of sorbent make-up (mol s−1)
ECO2, Carb Carbonator CO2 capture efficiency (mol mol−1) NCaO, Loop Molar flow of circulating CaO (mol s−1)
NCaO, Calc Molar amount of CaO in the calciner (mol) msolid Specific solid circulation rate (kg m−2 s−1)
NCaO, Carb Molar amount of CaO in the carbonator (mol) γi Mass fraction of component i (kg kg−1)
NCaO, active Molar amount of active CaO (mol) τCaO Residence time of CaO in the Carbonator (h)
NCycle Number of calcination and carbonation cycles (1) τCarb Residence time of CO2 in the carbonator (h)
Nage Conducted Number of calcination and carbonation cycles τactive Active space time (s)
(1) φ Gas-solid contact factor (1)
R Universal gas constant (J mol−1 K−1) dX/dt Carbonation reaction rate (mol s−1)
T Temperature (K) fCarb Fractional carbonation (mol mol−1)
TCalc Calcination temperature (∘C) factive Active fraction of particles (1)
TCarb Carbonation temperature (∘C) h Reactor height (m)
Ws, Calc Solid inventory of the calciner (kg m−2) ks Sorbent carbonation rate coefficient (s−1)
Ws, Carb Solid inventory of the carbonator (kg m−2) nLR Looping ratio (mol mol−1)
Ws Solid reactor inventory (kg m−2) nMR Make-up ratio (mol mol−1)
XCaCO3, Calc Carbonation degree of calciner sorbent (mol mol−1) pCO2, eq. Equilibrium CO2 partial pressure (bar)
XCaCO3, Carb Carbonation degree of carbonator sorbent (mol mol−1) t Time (h)
XCaCO3 Carbonation degree (mol mol−1) t∗ Critical reaction time (s)
Xavg Average sorbent CO2 carrying capacity (mol mol−1) xIL Integration level (mol mol−1)
Δp Pressure loss (mbar) yCO2, Carb, in CO2 concentration at carbonator inlet (m3 m−3)
yCO2 Logarithmic average CO2 concentration (m3 m−3) yCO2, Carb, out CO2 concentration at carbonator exit (m3 m−3)
NCO2, Carb, in Molar flow of CO2 entering the carbonator (mol s−1) yCO2, eq Equilibrium CO2 concentration (m3 m−3)
NCO2, Carb, out Molar flow of CO2 exiting the carbonator (mol s−1) yO2, Calc, in O2 concentration at calciner inlet (m3 m−3)
NCa, CaL Molar flow of calcium fed to the calcium looping system

CaO(s) + CO2(g) CaCO3(s) (1) 2. Calcium Looping CO2 capture from cement plants

The calcium looping technology is considered especially suitable for This work focuses on the carbonation reactor of the so called tail-
application in the cement sector, since both the clinker manufacturing end calcium looping option for cement plants in which the calcium
process and the calcium looping CO2 capture process share a common looping CO2 capture process is installed downstream of the cement
feedstock (i.e. CaCO3). This enables synergies such as sorbent re­ plant's pre-heater tower (Fig. 1). Another work focusing on the calciner
utilisation, increased make-up rates to the calcium looping system or can be found elsewhere [27]. This integration option uses limestone, or
energy integration. Several integration possibilities of calcium looping a limestone-rich raw meal fraction, as sorbent to run the calcium
CO2 capture into the cement plant have been developed and mostly looping CO2 capture process. Purged sorbent (i.e. CaO) is then re-uti­
assessed by process simulation. These include (i) an indirect, partial lised in the upstream cement plant replacing a fraction of the raw meal's
decarbonisation of the cement plant using calcined calcium looping limestone or calcium carbonate, respectively. Hence, the flue gas to be
purge (i.e. CaO) from a nearby power plant with calcium looping CO2 decarbonized by the calcium looping carbonator consists of (i) the ce­
capture as feedstock substitute for the clinker production [7,8], (ii) a ment kiln's combustion flue gas, (ii) the pre-calciner's combustion flue
tail-end calcium looping option using fluidised bed reactors to capture gas as well as (iii) the CO2 from raw meal calcination, (iv) reduced by
the cement plant's CO2 after its suspension pre-heater [9–11] and (iii) the share of calcium carbonate being already calcined by the calcium
an integrated calcium looping option using entrained flow reactors to looping process. The possibility to reuse calcium looping purge in the
capture the CO2 produced by clinker burning in the carbonator while cement plant allows the operation at high make-up rates. This influ­
CO2 from raw meal calcination is captured by oxy-fuel calcination in ences the calcium looping operation parameters such as sorbent ac­
the calcium looping calciner [11–13]. Since the tail-end integration tivity, and consequently, the required reactor solid inventory or sorbent
option is installed downstream of the cement plants pre-heater tower circulation rate between the calciner and carbonator, as well as the
such an option is easy to retrofit and there is only minor interference calciner and carbonator operation conditions. However, depending on
with the clinker burning process [14]. Due to enhanced fuel utilisation the cement plant's boundary conditions, primarily the available raw
the ash load in the cement kiln increases but the changes are compar­ meal composition (i.e. limestone to marl ratio of the cement plant's raw
able to the well-known behaviour after switching to high-ash fuels [14]. meal), operation with low make-up rates similar to power plant appli­
Apart from this, cement clinker has been successfully burned using cation may be favoured. The term integration level (xIL) is used to de­
calcium looping purge as limestone substitute by various research scribe the sorbent make-up fed to the calcium looping system with re­
groups [15–18] highlighting the suitability of calcium looping purge as spect to the cement clinker production. It is defined as molar amount of
clinker feedstock. Due to the inline installation of the integrated en­ calcium (NCa, CaL ) that is fed to the calcium looping subsystem (i.e.
trained flow option, such an integration is more complex to retrofit to sorbent make-up) and fed back to the cement plant, with respect to the
an existing cement plant but yield an improved performance due to a amount of calcium that is required to produce clinker in the upstream
favourable energy integration. CO2 capture cost in terms of primary cement plant (NCa, Clinker ).
energy demand (SPECCA) are estimated to range from 3.76 MJ kg−1 to
4.42 MJ kg−1 for the tail-end configuration depending on the integra­ NCa, CaL
xIL =
tion level and electricity scenario, whereas the integrated entrained NCa, Clinker (2)
flow configuration yield CO2 capture costs of 3.2 MJ kg−1 [11].
The integration level indirectly determines the make-up ratio of the
calcium looping system. With increasing integration level the CO2 load

2
M. Hornberger, et al. Fuel Processing Technology 210 (2020) 106557

amount of CaO with respect to incoming amount of CO2 (nLR), in­


sufficient space time (τCarb) or by the equilibrium CO2 partial pressure
(yCO2, eq). The latter defines the minimal achievable CO2 concentration
in the carbonator for a given temperature. Thus, the extent to which
CO2 is able to react with CaO. The equilibrium partial pressure was
determined from the FactPS [19] database, and can be described in
analogy to Silcox et al. [20] by Eq. (4) in the range from 500 ∘C to 1000

C with the universal gas constant (R) and an activation energy (EA) of
169.5 kJ mol−1.
EA
pCO2, eq. = 4.0009 107bar exp
RT (4)

The space time is defined as molar amount of CaO in the carbonator


in relation to the incoming flow of CO2. Hence, it is a measure of the
available reaction time of CO2 and CaO in the carbonator (Eq. (5)).
NCaO, Carb
Carb =
NCO2, Carb, in (5)

Additionally, the looping ratio is defined as molar amount of CaO


entering the carbonator in relation to the incoming amount of CO2.
Hence, it represents the amount of CaO in the carbonator that can react
with CO2 (Eq. (6)). However, since only a fraction of the circulating
CaO actually binds CO2 (Xavg), the active looping ratio is used to de­
scribe the CO2 capture potential of the circulating flow of CaO with
respect to the incoming amount of CO2 (Eq. (7)).

NCaO, Loop
nLR =
NCO2, Carb, in (6)

NCaO, Loop Xavg


Fig. 1. Tail-end calcium looping integration option for CO2 capture from ce­ nLR, active =
NCO2, Carb, in (7)
ment plants.
During the repetitive cycles the sorbent's ability to bind CO2 decays
attributed to the calcination in the cement plant's pre-calciner de­ exponentially towards an asymptotic threshold. Therefore, spent sor­
creases, while the fuel consumption of the rotary kiln remains constant. bent must continuously be purged and replaced by fresh sorbent make-
Theoretically, make-up ratios up to 4 mol mol−1 are possible assuming up. The molar amount of make-up fed to the calcium looping system
that (i) all CaCO3 is feed to the calcium looping system, that (ii) the with regard to the molar flow of CO2 entering the carbonator is referred
calcium looping purge is exiting fully calcined, and that (iii) the sen­ to as make-up ratio (nMR). This parameter affects primarily the sorbent's
sible heat of the kiln flue gas is sufficent to preheat the calcined raw CO2 carrying capacity and subsequently the required looping ratio to
meal. However, the reheating of cooled calcium looping purge is ex­ achieve a target CO2 capture efficiency.
pected to require additional energy provided by the fuel combustion. NCaCO3,0
Hence, the CO2 load towards the calcium looping carbonator as well as nMR =
NCO2, Carb, in (8)
the make-up ratio differs depending on the integration level, the carbon
intensity of the fuel burned and the additional energy required to reheat The make-up ratio of the tail-end calcium looping cement plant
the calcium looping purge. The additional energy required for purge integration is influenced by the integration level in two ways. Firstly
preheating in the cement plant's pre-calciner can be compensated by and directly, by the amount of CaCO3 (i.e. sorbent make-up) fed to­
means of energy recuperation from purge cooling. De Lena et al. [10] wards the calcium looping subsystem. Secondly, by the reduction of
assessed the tail-end calcium looping CO2 capture option using process CO2 emitted from the cement plant and fed to the calcium looping
simulation for five different integration levels (i.e. 15%, 20%, 25%, carbonator, since CO2 originating from the share of CaCO3 that is fed to
50% and 80%) reporting make-up ratios of 0.11 mol mol−1, 0.16 mol the calcium looping calciner is immediately captured by oxy-fuel cal­
mol−1, 0.21 mol mol−1, 0.60 mol mol−1 and 1.84 mol mol−1, utilising cination. Thus, the make-up ratio increases disproportionately with
coal as fuel in the cement plant. Operation with both high and relatively increasing integration level.
low make-up rates (i.e. integration levels) were experimentally assessed
at University of Stuttgart's 200 kW calcium looping pilot plant. 4. Experimental section

3. Theoretical background The experimental investigation of calcium looping CO2 capture from
cement plants presented in this work were conducted at University of
A crucial parameter for every CO2 capture technology is its CO2 Stuttgart's 200 kW fluidised bed pilot plant. The pilot plant consists of
capture performance. The CO2 capture efficiency of the calcium looping three refractory lined fluidised bed reactors that can either be inter­
carbonator is defined as the absorbed flow of CO2 divided by the in­ connected to a CFB-CFB configuration (Fig. 2a) or to a BFB-CFB con­
coming flow of CO2 (Eq. (3)). figuration (Fig. 2b). In both configurations, the calcination reactor is a
NCO2, Carb, in NCO2, Carb, out circulating fluidised bed reactor while, depending on the configuration,
ECO2, Carb = either the CFB carbonator or the BFB carbonator is deployed. Both CFB
NCO2, Carb, in (3)
reactors are 10.8 m high and have an inner diameter of approx.
For the calcium looping process the amount of CO2 that can be 200 mm. The calciner's inner diameter increases gradually from
captured in the carbonator can either be limited by the incoming 120 mm to 210 mm in the lower third of the reactor. Contrary, the CFB

3
M. Hornberger, et al. Fuel Processing Technology 210 (2020) 106557

Volume flows, gas composition at the calciner outlet as well as carbo­


nator inlet and outlet, pressure and temperature, and parameters de­
rived from these values are continuously monitored, calculated and
recorded, whereas solid samples are collected periodically. The circu­
lation rate between the reactors is constantly monitored by microwave
sensors (CFB-CFB configuration) or the conveyor screw's rotation speed
(BFB-CFB configuration) and manually measured and verified by means
of solid accumulation in a dedicated measuring section. The reactors'
solid inventories are calculated based on the pressure drop over each
reactor part with its respective cross-section and summed up. Subse­
quently, the molar calcium inventory was determined based on the
chemical analysis of the collected bed samples. The reference tem­
perature of the reactors were calculated by integrating and averaging
the temperature measurements along the reactor height, since the
thermocouples are not installed equidistantly along the reactor height.
As the oxidation gas is fed to the calciner in a staged manner, the in­
tegration starts after all oxidant is injected to the calciner (i.e. after the
third stage or 4 m, respectively). For the CFB carbonator the tempera­
ture is integrated from the gas distributor to the reactor outlet, whereas
for the BFB carbonator the temperature is integrated along the carbo­
nator bed. The gas composition is measured continuously by non-dis­
persive infrared spectroscopy (CO, CO2, SO2, NOx), paramagnetism (O2)
and impact jet psychrometry (H2O). In this work, a broad range of
operation conditions representing those expected for calcium CO2
capture from cement plant integration were investigated using Co­
lumbian hard coal as fuel and a high purity limestone from western
Germany as sorbent. Their respective chemical composition are sum­
marized in Tables 1 and 2, whereas the investigated operation condi­
tions are summarized in Table 3. Initially, the effect of high integration
levels on the performance of the calcination and carbonation reactor
was assessed at different calcination and carbonation temperatures with
a flue gas CO2 concentration of 0.15 m3 m−3. The maximum feeding
rate of limestone was limited to 50 kg h−1 by the feeders rotational
speed which limited the make-up rate to approx. 0.6 mol mol−1. Hence,
the BFB configuration was employed to assess even higher make-up
Fig. 2. University of Stuttgart's 200 kW calcium looping pilot plant at the ratios of 0.9 mol mol−1. Low integration levels were assessed using the
Institute of Combustion and Power Plant Technology (IFK) - (a) CFB-CFB con­
CFB-CFB configuration by a variety of make-up rates and CO2 con­
figuration, (b) BFB-CFB configuration.
centrations at a carbonation temperature of 645 ∘C, known to be ben­
eficial for less active sorbent, and a calcination temperature of approx.
carbonator holds an increased inner diameter of 330 mm at the bottom 920 ∘C.
that is reduced to 220 mm at its riser section. This reactor design allows
for longer residence time in the bottom section of the reactor (i.e. its
5. Results and discussion
dense bed region). The BFB carbonator has a height of 6 m and an inner
diameter of 330 mm. Subsequently to each fluidised bed riser a cyclone
The experiments presented in this work were obtained during four
separates the entrained solids from the gas. Additionally, a protective
experimental campaigns. Each experiment was operated for at least one
cyclone is installed before each bag filter to separate finer particles and
hour (but in general 2 h), after steady state conditions were reached.
to protect the bag filter in case of malfunction. Solids separated by the
The transition from one steady-state point to another takes a few hours
cyclone are directed back to the reactors via loop seals, whereas the
depending on the severity of the imposed changes. Throughout the
solids separated by the protective cyclone are collected in a bin. For
presented experimental points the pilot facility was operated with a
each reactor, temperature and pressure are measured along the reactor
solid inventory of approx. 80 kg or 970 kg m−2 in the carbonator and
height, the solid circulation lines and the exhaust gas lines. In case of
approx. 20 kg or 450 kg m−2 in the calciner.
CFB-CFB configuration, solids separated by the cyclone are either re­
cycled internally or directed to the other reactor. The respective share
of external and internal circulation is controlled via a cone valve in the 5.1. Experimental operation
loop seal. This design allows independent investigation of circulation
rate from solid entrainment. However, for the BFB-CFB configuration, A time trend of an experiment held for 6 h is presented in Fig. 3. The
solids are fed from the calciner to the carbonator via a screw feeder and calciner was operated with a specific solid inventory of approx. 430 kg
cycled back to the calciner via a loop seal at the bottom of the reactor. m−2 and an average riser temperature of 933 ∘C, whereas the
Solid circulation is solely controlled by the screw feeder's rotational
speed. Solids (i.e. limestone and coal) are fed into the calciner by means Table 1
of a gravimetrically controlled screw feeders. For oxy-fuel operation, Chemical composition of utilised sorbent.
CO2 rich exhaust gas is recycled and mixed with oxygen for fuel com­ γCaO γCO2 γOthers
bustion. The gas mixture can be fed in three stages to control the
combustion performance in the calciner. However, the carbonator is fed kg kg−1, wf
with a synthetic flue gas generated by mixing air, steam and CO2 that
Limestone 0.5465 0.4357 0.0178
can be electrically pre-heated in case of the CFB-CFB configuration.

4
M. Hornberger, et al. Fuel Processing Technology 210 (2020) 106557

Table 2 5.2. Hydrodynamic


Chemical composition of utilised hard coal.
Coal γC γH γO γN γS γash γH2O A stable hydrodynamic of the dual fluidised bed system is essential
for the operation of a calcium looping system. Both carbonator reactor
kg kg−1, waf kg kg−1, wf kg kg−1, ad concepts were successfully operated throughout multiple experimental
campaigns. Hydrodynamic profiles of the respective reactor config­
Columbian I 0.803 0.049 0.123 0.019 0.006 0.096 0.074
Columbian II 0.776 0.053 0.144 0.016 0.011 0.091 0.074 uration (i.e. BFB-CFB and CFB-CFB) are depicted in Fig. 4. The pressure
gradient of the reactors is depicted as solid lines, whereas internal
ad: air dried as used in experiments circulation and transfer between the reactors is depicted as dashed lines
and densely dotted lines, respectively. The pressure gradient of the
calciner shows a smooth transition from the bottom section into a linear
Table 3 gradient at the calciner's riser section for both configurations. The
Experimental operation conditions for tail-end CaL CO2 capture from cement calciner's solid inventory is slightly increased when the BFB carbonator
plants. is employed to provide sufficient ‘counter pressure’ to enable the car­
Parameter Variable Value/Range Unit bonator to hold the reference carbonator mass. The BFB carbonator
shows a typical pressure gradient in which solids are accumulated at the
Carbonator temperature TCarb 585–700 ∘
C
reactor bottom and form a dense region with minor entrainment of
Carbonator solid inventory Ws, Carb 680–1270 kg m−2
Carbonator CO2 inlet concentration yCO2, Carb, in 0.13–0.35 m3 m−3 solids into the freeboard region. The pressure gradient of the CFB car­
Carbonator CO2 capture ECO2, Carb 0.45–0.98 mol mol−1 bonator shows an accumulation of solid in its enlarged bottom section
Calciner temperature TCalc 890–930 ∘
C and a sharp transition into a linear gradient at its riser section when the
Calciner solid inventory Ws, Calc 290–840 kg m−2 carbonator's cross-section is reduced. This design was originally chosen
Calciner O2 inlet concentration yO2, Calc, in 0.42–0.65 m3 m−3
Make-up ratio nMR 0.08–0.9 mol mol−1
to reduce the gas velocity for a given flue gas volume flow. This allows
Looping ratio nLR 4–14 mol mol−1 for longer residence times in the dense region and, thus, enhance CO2
capture. However, this design has an adverse effect when operating the
carbonator with high CO2 concentrations. A significant reduction of
carbonator was operated with a solid inventory of 680 kg m−2 and a fluidisation gas, or gas velocity respectively, occur at high CO2 capture
bed temperature of 630 ∘C. Within the 6 h time frame the circulation efficiencies in combination with high CO2 concentrations. This reduc­
rate was measured four times yielding consistent circulation rates of tion can be as high as 30% for CO2 capture from cement plants. This
390 kg h−1 or 1.35 kg m−2 s−1 respectively. The circulation rate was phenomena was intensified due to the CFB carbonator's enlarged
referred to the carbonator's reference cross-section, which was de­ bottom cross-section further reducing the gas velocity and resulted in
termined based on the solid distribution in the carbonator. Despite the insufficient solid entrainment from the carbonator bed. Proper solid
high CO2 carrying capacity of the sorbent, the CO2 capture averaged entrainment was ensured by increasing the total volume flow fed to the
around 80%. The CO2 capture was limited by insufficient circulation of carbonator. Consequently, the CO2 load towards the carbonator was
sorbent. While the circulating sorbent flow was able to absorb 0.76 mol increased which resulted in reduced CO2 capture efficiencies at given
h−1 of CO2, the CO2 feed into the carbonator was 0.87 mol h−1. operation conditions (i.e. circulation rate, solid inventory, carbonation
temperature). The reduction of volume flow and gas velocity of up to
30% by CO2 capture can easily be addressed when designing a

Fig. 3. Time trend of an experiment operated for 6 h. Upper graph: Temperature (T) and specific solid inventory (WS) of the calciner (Calc) and carboantor (Carb);
lower graph: Carbonator CO2 capture efficiency (ECO2, Carb), carbonator CO2 outlet concentration (yCO2, Carb, out), solid carbonation degree (XCaCO3,□) and specific solid
circulation rate (ms ).

5
M. Hornberger, et al. Fuel Processing Technology 210 (2020) 106557

Fig. 4. Hydrodynamic profile depicted as reactor height (h) against the pressure loss (Δp) for the BFB-CFB configuration (top) and the CFB-CFB configuration
(bottom).

commercial carbonator unit. Moreover, a CFB carbonator design is most


likely to be employed for cement plant application since it allows lager
volume flows per cross-section and has a superior gas solid contact over
the BFB carbonator.

5.3. CO2 capture performance

The operation and design of the calcium looping carbonator de­


pends on the sorbent activity and the CO2 loading, or in case of cement
plant integration on the integration level of the calcium looping system
respectively. For high integration levels (i.e. make-up rates) the CO2
loading towards the carbonator is significantly reduced since a large
share of the cement plant's CO2 emissions bound in the raw meal's
carbonates are captured by oxy-fuel calcination in the calcium looping
calciner. Simultaneously, the sorbent activity increases. The CO2 cap­
ture efficiencies achieved during the investigation of high integration
levels are presented in Fig. 5. For these experiments, the make-up ratio
was 0.6 mol mol−1 (circles) and 0.9 mol mol−1 (squares) with a CO2
concentration of 0.15 m3 m−3. It can clearly be seen that the CO2
capture efficiency was limited by the achievable equilibrium CO2 con­
centration or equilibrium CO2 capture, respectively. For such operation
Fig. 5. CO2 capture efficiency (ECO2, Carb) vs. carbonation temperature (TCarb)
conditions, capture efficiencies up to 98% were achieved in the car­
for selected experiments representing high integration levels (i.e. make-up rates
bonator at temperatures slightly below 600 ∘C. The good capture per­ of 50 kg h−1 and a reduced CO2 concentration of 0.15 m3 m−3) using the CFB
formance can be ascribe to the sorbent's high CO2 carrying capacity, carbonator (nMR=0.6 mol mol−1) and the BFB carbonator
resulting in an excess of unreacted CaO for the present circulation rates. (nMR=0.9 mol mol−1).
Thus, allowing to capture CO2 as low as the equilibrium CO2 con­
centration at the carbonator outlet. The sorbent CO2 carrying capacity
while the sorbent activity decreases towards residual activity. The
ranged from 0.2 mol mol−1 up to 0.37 mol mol−1 corresponding to an
lower sorbent activity can be counteracted by increasing sorbent cir­
active looping ratio (nLR ⋅ Xavg) between 0.9 mol mol−1 and 5 mol
culation or looping ratio, respectively, to achieve a target CO2 capture.
mol−1. The high capture rates at lower temperature may also be fa­
Fig. 6 presents the achieved CO2 capture efficiency in the carbonator
voured by the comparatively large surface area of the low cycled sor­
against the active looping ratio for cases which represent low integra­
bent. The accessible surface area enhances the CO2 capture especially at
tion levels at a carbonation temperature of approx. 645 ∘C categorized
low CO2 partial pressures when the driving force diminishes. Similarly,
according to their sorbent activity. Supplementary, the experiments
Lu et al. [21] yielded improved CO2 capture performance for low cycled
investigating high integration levels are depicted (squares) in Fig. 6. A
sorbent at temperature around 600 ∘C which was attributed to a higher
linear increase of CO2 capture efficiency with the active looping ratio is
accessible surface area of the sorbent.
evident for an active looping ratio below 1 mol mol−1, indicating that
Contrarily, if operated at low integration levels, the CO2 load to the
the CO2 capture is limited by the amount of active calcium being fed to
carbonator increases resulting in CO2 concentrations up to 0.33 m3 m−3

6
M. Hornberger, et al. Fuel Processing Technology 210 (2020) 106557

(circles) fits the model fairly well, a strong deviation from the model
line (white diamonds) can be observed for lower cycled sorbent (i.e.
high make-up rates). The high deviation of these samples can be as­
cribed to partial carbonation of the circulating solids. Due to a surplus
amount of active CaO available in the carbonator the sorbent was not
able to be fully carbonised. This phenomenon occurred at high make-up
rates for BFB operation as well as for CFB operation at higher carbo­
nation temperatures (i.e. above 675 ∘C). At higher carbonation tem­
peratures the absorbable amount of CO2 was reduced by the achievable
equilibrium CO2 concentration, whereas in case of the BFB operation,
the sorbent's CO2 uptake was limited by the significantly reduced
amount of CO2 being fed to the carbonator. As a consequence, such
samples undergo a reduced number of calcination and carbonation
cycles resulting in a younger sorbent age. Sample which had a lower
fractional carbonation degree (Eq. (10)) below 0.5 mol mol−1 were
corrected using Eq. (11). Unsurprisingly, such sample coincide with
experiments operated with an active looping ratio above 1.5 mol
mol−1. The term fractional carbonation degree describes the achieved
carbonation conversion (XCaCO3, Carb − XCaCO3, Calc) with respect to the
achievable carbonation degree (i.e. CO2 carrying capacity (Xavg)). Cor­
recting the sorbent age with the sorbent's fractional conversion (fCarb)
yield reasonable agreement with the predicted CO2 carrying capacity of
Fig. 6. CO2 capture efficiency (ECO2, Carb) vs. active looping ratio (nLR ⋅ Xavg) the collected samples (black diamonds).
categorized by sorbent activity for experiments representing low integration
levels (i.e. low make-up rates and high CO2 concentrations) at carbonation XCaCO3, Carb XCaCO3, Calc
fCarb =
temperature of approx. 645 ∘C. Additionally, high integration cases (i.e. Xavg ≥ Xavg (10)
0.2) are depicted.
Nage = NCycle fCarb (11)
the carbonator. Hence, the CO2 capture can be adjusted by means of the
At this point it should be highlighted that such a state, in which
sorbent circulation rate. Optimising the looping ratio to achieve a target
solids are not fully carbonated in the carbonator, represent an in­
CO2 capture efficiency of 90% in the carbonator for low integration
efficient state of operation. The unused but circulated sorbent increases
cases required a looping ratio of 7 mol mol−1 when operating the pilot
heat losses and consequently reduces the overall efficiency of the cal­
plant with a sorbent activity of 0.13 mol mol−1. For active looping
cium looping process. However, a well designed, commercial calcium
ratios larger than 1 mol mol−1 a surplus amount of calcium is being
looping system can easily avoid operational cases like this.
circulated between the reactors. As a consequence, the CO2 capture is
limited by the achievable equilibrium CO2 concentration. Such opera­
tion conditions were observed for the investigation of the high in­ 5.5. Carbonator model/active space time approach
tegration levels, due to the significantly increase sorbent activity
maintaining similar circulation rates. Operation with both high make- The active space time approach developed by Charitos [23,24] re­
up rates and high sorbent activity, as well as low make-up rates and presents a simplified carbonator CO2 balance that can be used for a
residual sorbent activity are able to capture more than 90% of the in­ basic design of the carbonator. It was applied with satisfactory
coming CO2 in the carbonator, which is usually considered the target
CO2 capture performance for post-combustion CCS processes. It must be
highlighted that CO2 capture from oxy-fuel calcination of the cement
plant's raw meal is not included in these capture rates. Thus, the overall
CO2 capture of a cement plant with calcium looping CO2 capture will be
reasonably higher.

5.4. Sorbent performance

As expected, the sorbent activity decreases with decreasing in­


tegration level. The screening of the low integration level yields CO2
carrying capacities between 0.11 mol mol−1 and 0.135 mol mol−1,
whereas the high integration cases yield sorbent activities above
0.2 mol mol−1. Fig. 7 presents the sorbent's CO2 carrying capacity
determined from collected solid samples using thermogravimetric
analysis (TGA) against the number of calcination and carbonation cy­
cles for all conducted experiments. The cycle number was determined
based on the cycle frequency and the sorbents residence time in the
calcium looping system using Eq. (9).
NCaO, Loop NCaO, Carb + NCaO, Calc
NCycle =
NCaO, Carb + NCaO, Calc NCaCO3,0 (9)
The CO2 carrying capacity of the used limestone determined by TGA Fig. 7. Sorbent CO2 carrying capacity (Xavg) vs. calcination and carbonation
analysis and fitted to the model of Grasa et al. [22] is depicted as da­ cycles (NCycle) for operation representing high and low integration levels be­
shed line. While the CO2 carrying capacity of higher cycled sorbent tween the calcium looping CO2 capture process and the cement plant.

7
M. Hornberger, et al. Fuel Processing Technology 210 (2020) 106557

agreement to calcium looping operation from power plants up to semi


industrial scale [4,6] and at small pilot scale for operation conditions
representing CO2 capture from cement plants [25]. For defined opera­
tion conditions (i.e. carbonator temperature, CO2 concentration, molar
CaO inventory, solid circulation rate) and a given sorbent or limestone
(i.e. reaction rate, CO2 carrying capacity) the captured amount of CO2
and hence the CO2 capture efficiency can be calculated. The amount of
CO2 removed from the gas phase equals the CO2 absorbed by the CaO
particles in the carbonator bed. The removal of CO2 from the gas phase
can be described by the left side of Eq. (12), while the solid CO2 uptake
or carbonation is described by the right side of Eq. (12). The approach
assumes that the sorbent reacts with a constant reaction rate (dX/dt),
which is characteristic for each sorbent, until it reaches its CO2 carrying
capacity (Xavg). Sorbent that holds a lower residence time than the
critical reaction time (t∗) is not fully loaded and therefore participates
at the CO2 uptake. Assuming a CSTR residence time distribution for the
solids in the carbonator, the fraction of active or reacting sorbent (factive)
can be calculated by Eq. (13). Therefore, the amount of active CaO
(NCaO, active) can be expressed as the product of NCaO, Carb and factive.
dX
NCO2, Carb, in ECO2, Carb = NCaO, active
dt Carb (12)
Fig. 8. Normalized CO2 capture efficiency (ECO2, Carb/ECO2, Carb, eq) vs. active
t space time (τactive) for operation conditions representing calcium looping CO2
factive = 1 exp capture from cement plants using the CFB carbonator.
CaO (13)
The reaction rate can be expressed as a function of the driving force the conducted experiments emphasises that this basic carbonator design
of the CO2 partial pressure towards the equilibrium CO2 partial pressure approach is also valid for higher CO2 concentrations present for calcium
( yCO2 yCO2, eq ), the average sorbent activity and the reaction rate (φks) looping CO2 capture from cement plants.
(Eq. (14)). ks represents the sorbent's characteristic reaction rate and φ
a fitting factor for the gas-solid contact [23,26]. Consequently, the CO2
capture efficiency can be described by Eq. (15). 6. Conclusion

dX
= ks Xavg (yCO2 yCO2, eq ) Operation conditions that are anticipated for calcium looping CO2
dt Carb (14) capture from cement plant were experimentally investigated at
NCaO, Carb factive Xavg University of Stuttgart's 200 kW calcium looping pilot plant. As ex­
ECO2,Carb = ks (y¯CO2 yCO2, eq ) pected, operation with high make-up rates strongly increases the sor­
NCO2, Carb, in (15)
bent's CO2 carrying capacity as high as 0.38 mol mol−1, yielding CO2
Parameters that can directly or indirectly be influenced by means of capture efficiencies up to 98% in the carbonator not including the large
design such as sorbent activity (through sorbent make-up), circulation share of CO2 captured by oxy-fuel calcination of the cement raw meal's
rate and solid inventory are combined with respect to the molar flow of limestone. The experiments investigating low integration levels yielded
CO2 entering the carbonator in the so called active space time (τactive). reduced sorbent activity that can be compensated by increased sorbent
circulation to achieve a target CO2 capture. The basic design criteria for
ECO2 = active ks (yCO2 yCO2, eq ) (16) calcium looping CO2 capture known from power plant investigation
The CO2 capture efficiency of the CFB carbonator experiments is (i.e. active space time approach) proved to be also valid for calcium
plotted against the active space time in Fig. 8. The BFB experiments are looping systems operating at cement specific operation conditions.
not presented since the figure's readability would suffer severely due to Furthermore, the reduction of fluidisation gas and gas velocity up to
the significantly increased active space times of the BFB carbonator 30% by CO2 capture needs to be considered when designing a carbo­
experiments ranging from 150 s to 450 s. The two depicted model lines nator that is operated with high volume fractions of CO2 to ensure
represent the lower and the upper limit of the CO2 concentration that proper solid entrainment.
are anticipated for calcium looping CO2 capture from cement plants
with high and low integration levels. Hence, all conducted experiments CRediT authorship contribution statement
should lie between the two model lines. Considering uncertainties in
the determination of circulation rate, carbonation degree and average 1. M.Sc. Matthias Hornberger: Conceptualization, Methodology,
sorbent CO2 carrying capacity and consequently the active space time, Validation, Investigation, Data Curation, Writing – Original Draft,
the experiments show a very good agreement with the prediction of the Visualization
active space time model. For the conducted CFB carbonator experi­ 2. M. Sc. Joseba Moreno: Review & Editing
ments an apparent reaction rate (ksφ) of 0.30 s−1 with a gas solid 3. M. Sc. Max Schmid: Review & Editing
contact factor (φ) of 0.92 was determined, whereas the BFB carbonator 4. Univ.-Prof. Dr. techn. Günter Scheffknecht: Project Administration,
yielded an apparent reaction rate (ksφ) of 0.09 s−1 with a gas solid Supervision, Review & Editing
contact factor (φ) of 0.24. The distinctive poor gas solid contact of the
BFB carbonator results from bubble formation allowing the gas to by-
pass the sorbent bed. Contrary, the CFB carbonator shows very good gas Declaration of competing interest
solid contacting due to the operation in the fast fluidisation regime. As a
consequence, CFB reactors are more likely to be employed in calcium The authors declare that they have no known competing financial
looping systems operating with constant volume flows such as in ce­ interests or personal relationships that could appeared to influence the
ment plants. The agreement between the active space time model and work reported in this paper.

8
M. Hornberger, et al. Fuel Processing Technology 210 (2020) 106557

Acknowledgements Procedia 61 (2014) 500–503, https://doi.org/10.1016/j.egypro.2014.11.1158.


[13] M. Spinelli, I. Martínez, M.C. Romano, One-dimensional model of entrained-flow
carbonator for CO2 capture in cement kilns by Calcium looping process, Chem. Eng.
This project has received funding from the European Union's Sci. 191 (2018) 100–114, https://doi.org/10.1016/j.ces.2018.06.051.
Horizon 2020 research and innovation programme under grant agree­ [14] H. Hoppe, V. Hoenig, J. Ruppert, D. Berstad, M. Romano, D. Sutter, M. & Voldsund,
ment no 641185 (CEMCAP). D4.5 Retrofitability study for CO2 capture technologies in cement plants., Zenodo
(2018).
A special thanks to Heiko Holz without whom the pilot plant could [15] C. Dean, T. Hills, N. Florin, D. Dugwell, P.S. Fennell, Integrating Calcium Looping
not be operated so smoothly as well as many thanks to the department CO2 Capture with the Manufacture of Cement, Energy Procedia 37 (2013)
for the assistance during the weeklong experimental campaigns. 7078–7090, https://doi.org/10.1016/j.egypro.2013.06.644.
[16] J. Ströhle, Volker Hönig, IGF - Carbonate Looping Schlussbericht: Entwicklung
eines Konzepts zur CO2-Abscheidung durch Carbgonate Looping mit Verwertung
References der Absorbentien in der Zementindustrie (2013). URL http://www.dvv.uni-
duisburg-essen.de/projekte/pdf/19_361ZN_Schlussbericht.pdf.
[17] A. Telesca, D. Calabrese, M. Marroccoli, M. Tomasulo, G.L. Valenti, G. Duelli,
[1] IEA, Technology Roadmap - Low-Carbon Transition in the Cement Industry (2018).
F. Montagnaro, Spent limestone sorbent from calcium looping cycle as a raw ma­
URL https://www.iea.org/reports/technology-roadmap-low-carbon-transition-in-
terial for the cement industry, Fuel 118 (2014) 202–205, https://doi.org/10.1016/
the-cement-industry.
j.fuel.2013.10.060.
[2] F. Schorcht, I. Kourti, B. M. Scalet, S. Roudier, L. Delgado Sancho, Best available
[18] M. Erans, M. Jeremias, L. Zheng, J.G. Yao, J. Blamey, V. Manovic, P.S. Fennell,
techniques (BAT) reference document for the production of cement, lime and
E.J. Anthony, Pilot testing of enhanced sorbents for calcium looping with cement
magnesium oxide: Industrial Emissions Directive 2010/75/EU (integrated pollution
production, Appl. Energy 225 (2018) 392–401, https://doi.org/10.1016/j.
prevention and control), Vol. 26129 of EUR, Scientific and technical research series,
apenergy.2018.05.039.
Publications Office, Luxembourg, 2013.
[19] C.W. Bale, E. Bélisle, P. Chartrand, S.A. Decterov, G. Eriksson, A.E. Gheribi, K. Hack,
[3] T. Shimizu, T. Hirama, H. Hosoda, K. Kitano, M. Inagaki, K. Tejima, A Twin Fluid-
I.-H. Jung, Y.-B. Kang, J. Melançon, A.D. Pelton, S. Petersen, C. Robelin, J. Sangster,
Bed Reactor for Removal of CO2 from Combustion Processes, Chem. Eng. Res. Des.
P. Spencer, M.-A. van Ende, FactSage thermochemical software and databases,
77 (1) (1999) 62–68, https://doi.org/10.1205/026387699525882.
2010–2016, Calphad 54 (2016) 35–53, https://doi.org/10.1016/j.calphad.2016.05.
[4] B. Arias, M.E. Diego, A. Méndez, M. Alonso, J.C. Abanades, Calcium looping per­
002.
formance under extreme oxy-fuel combustion conditions in the calciner, Fuel 222
[20] Geoffrey D. Silcox, John C. Kramlich, W. David, Pershing, A mathematical model
(2018) 711–717, https://doi.org/10.1016/j.fuel.2018.02.163.
for the flash calcination of dispersed calcium carbonate and calcium hydroxide
[5] H. Dieter, A.R. Bidwe, G. Varela-Duelli, A. Charitos, C. Hawthorne, G. Scheffknecht,
particles, Industrial & Engineering Chemistry Research, 1989, pp. 155–160.
Development of the calcium looping CO2 capture technology from lab to pilot scale
[21] D.Y. Lu, R.W. Hughes, E.J. Anthony, Ca-based sorbent looping combustion for CO2
at IFK, University of Stuttgart, Fuel 127 (2014) 23–37, https://doi.org/10.1016/j.
capture in pilot-scale dual fluidized beds, Fuel Process. Technol. 89 (12) (2008)
fuel.2014.01.063.
1386–1395, https://doi.org/10.1016/j.fuproc.2008.06.011.
[6] J. Hilz, M. Helbig, M. Haaf, A. Daikeler, J. Ströhle, B. Epple, Long-term pilot testing
[22] S.G. Grasa, J.C. Abanades, CO2 Capture Capacity of CaO in Long Series of
of the carbonate looping process in 1 MWth scale, Fuel 210 (2017) 892–899,
Carbonation/Calcination Cycles, Ind. Eng. Chem. Res. 45 (26) (2006) 8846–8851,
https://doi.org/10.1016/j.fuel.2017.08.105.
https://doi.org/10.1021/ie0606946.
[7] L.M. Romeo, D. Catalina, P. Lisbona, Y. Lara, A. Martínez, Reduction of greenhouse
[23] A. Charitos, N. Rodríguez, C. Hawthorne, M. Alonso, M. Zieba, B. Arias,
gas emissions by integration of cement plants, power plants, and CO2 capture
G. Kopanakis, G. Scheffknecht, J.C. Abanades, Experimental Validation of the
systems, Greenhouse Gases: Science and Technology 1 (1) (2011) 72–82, https://
Calcium Looping CO2 Capture Process with two Circulating Fluidized Bed
doi.org/10.1002/ghg3.5.
Carbonator Reactors, Ind. Eng. Chem. Res. 50 (16) (2011) 9685–9695, https://doi.
[8] M.C. Romano, M. Spinelli, S. Campanari, S. Consonni, G. Cinti, M. Marchi,
org/10.1021/ie200579f.
E. Borgarello, The Calcium Looping Process for Low CO2 Emission Cement and
[24] A. Charitos, Experimental Characterization of the Calcium Looping Process for CO2
Power, Energy Procedia 37 (2013) 7091–7099, https://doi.org/10.1016/j.egypro.
Capture, Dissertation, University of Stuttgart, Stuttgart, Germany (2013). doi:10.
2013.06.645.
18419/opus-2174.
[9] N. Rodríguez, R. Murillo, J.C. Abanades, CO2 capture from cement plants using
[25] B. Arias, M. Alonso, C. Abanades, CO2 capture by calcium looping at relevant
oxyfired precalcination and/or calcium looping, Environmental science & tech­
conditions for cement plants: experimental testing in a 30 kW th pilot plant, Ind.
nology 46 (4) (2012) 2460–2466, https://doi.org/10.1021/es2030593.
Eng. Chem. Res. 56 (10) (2017) 2634–2640, https://doi.org/10.1021/acs.iecr.
[10] E. de Lena, M. Spinelli, I. Martínez, M. Gatti, R. Scaccabarozzi, G. Cinti,
6b04617.
M.C. Romano, Process integration study of tail-end Ca-Looping process for CO2
[26] N. Rodríguez, M. Alonso, J.C. Abanades, Experimental investigation of a circulating
capture in cement plants, International Journal of Greenhouse Gas Control 67
fluidized-bed reactor to capture CO2 with CaO, AICHE J. 57 (5) (2011) 1356–1366,
(2017) 71–92, https://doi.org/10.1016/j.ijggc.2017.10.005.
https://doi.org/10.1002/aic.12337.
[11] E. de Lena, M. Spinelli, M. Gatti, R. Scaccabarozzi, S. Campanari, S. Consonni,
[27] M. Hornberger, J. Moreno, M. Schmid, G. Scheffknecht. Experimental investigation
G. Cinti, M.C. Romano, Techno-economic analysis of calcium looping processes for
of the calcination reactor in a tail-end Calcium Looping configuration for CO2
low CO2 emission cement plants, International Journal of Greenhouse Gas Control
capture from cement plants. Fuel, In Press, https://doi.org/10.1016/j.fuel.2020.
82 (2019) 244–260, https://doi.org/10.1016/j.ijggc.2019.01.005.
118927.
[12] M.C. Romano, M. Spinelli, S. Campanari, S. Consonni, M. Marchi, N. Pimpinelli,
G. Cinti, The Calcium Looping Process for Low CO2 Emission Cement Plants, Energy

You might also like