You are on page 1of 9

Fuel 150 (2015) 269277

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

On steam hydration of CaO-based sorbent cycled for CO2 capture


John Blamey a, Vasilije Manovic b, Edward J. Anthony b, Denis R. Dugwell a, Paul S. Fennell a,
a
b

Department of Chemical Engineering, Imperial College London, South Kensington, London SW7 2AZ, UK
Centre for Combustion and CCS, Craneld University, Craneld, Bedfordshire MK43 0AL, UK

h i g h l i g h t s
 Steam hydration to increase reactivity of CaO for CO2 capture has been investigated.
 The reactivity and friability of reactivated sorbent was investigated.
 Carbonation extent was found to have a linear relationship with hydration extent.
 Hydration was found to have a strong relationship with porosity/cycling history.
 As a result, the optimal hydration frequency can be predicted.

a r t i c l e

i n f o

Article history:
Received 6 December 2014
Received in revised form 2 February 2015
Accepted 5 February 2015
Available online 20 February 2015
Keywords:
Calcium looping
Hydration
Fluidised bed
Carbon capture

a b s t r a c t
The reversible reaction of CaO with CO2 can be used for post- and pre-combustion capture of CO2; however, the reactivity of CaO particles is found to reduce upon repeated use. Hydration has been shown to be
an effective method of increasing the reactivity of (or reactivating) CaO to CO2 for CO2 capture. Here, a
lab-scale uidised bed reactor was used to investigate reactivation of sorbent using steam at two different hydration temperatures of 473 and 673 K. Prior to hydration, the sorbent was cycled at three different
calcination temperatures of 1123, 1173 and 1223 K; the carbonation temperature was kept constant at
973 K. Following hydration, the sorbent was either carbonated directly or carbonated indirectly via a
CaO intermediate. The hydration extent was found to decrease with increasing calcination temperature
before hydration and with increasing hydration temperature. The carbonation extent following hydration
was found to increase linearly with hydration extent depending on the method of carbonation with
direct carbonation resulting in higher conversions. Mass loss from the uidised bed was found to be higher for lower hydration temperatures and increased calcination temperatures before cycling. The hydration behaviour of sorbent was subsequently investigated using a TGA at three different steam
hydration temperatures of 483, 578 and 678 K. Material for the TGA tests was prepared using the lab-scale uidised bed reactor, with a calcination temperature of 1173 K and a carbonation temperature of
973 K. In this case, the number of cycles was varied from 0 to 13, in order to provide a wider range of
sorbent properties. Data from the TGA were used to project subsequent carbonation conversions and
relative increases in carrying capacity across hydration. The TGA hydration tests emphasise the importance of hydration temperature and prior cycling conditions and length on the increase in carrying capacity following hydration.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
The calcium looping (CaL) cycle [1] is a high temperature process that can be used to capture CO2 from dilute sources, e.g., ue
gases from coal power plants, to provide a concentrated stream

Corresponding author. Tel.: +44 (0) 20 7594 6637.


E-mail addresses: jblamey@imperial.ac.uk (J. Blamey), v.manovic@craneld.ac.
uk (V. Manovic), b.j.anthony@craneld.ac.uk (E.J. Anthony), d.dugwell@imperial.ac.
uk (D.R. Dugwell), p.fennell@imperial.ac.uk (P.S. Fennell).
http://dx.doi.org/10.1016/j.fuel.2015.02.026
0016-2361/ 2015 Elsevier Ltd. All rights reserved.

suitable for storage. The CaL cycle has recently been demonstrated
at pilot scale (11.7 MWth) [2,3]. The key reaction is the reversible
carbonation of calcium oxide (see (Rn.1)), which can be generated
easily and cheaply from limestone (calcium carbonate). A simplied typical proposed post-combustion process with an added
hydration/reactivation loop (with material either returned to the
carbonator or calciner) is shown in Fig. 1. CaCO3 is added to the calciner (typically 1193 K at 101 kPa), whereupon it calcines to form
CaO and CO2. The CO2 is pressurised and piped to a suitable storage
site and the CaO is returned to the carbonator. In the carbonator,

270

J. Blamey et al. / Fuel 150 (2015) 269277

the CaO is contacted with ue gas and reacts with any CO2 present
(typically this occurs around 923 K, at 101 kPa) to form CaCO3. The
CaCO3 is then recycled to the calciner and the CO2 lean gas can be
cooled and released to atmosphere:

CaOs CO2g
CaCO3s

DHhr 178 kJ=mol

CaOs H2 Og
CaOH2s

Rn:1

DHhr 109 kJ=mol

CaOH2s CO2g
CaCO3s H2 Og

Rn:2

DHhr 69 kJ=mol


Rn:3

An important aspect of CaL is that the CaO does not fully recarbonate in the carbonation stage and the recarbonation extent (or
carrying capacity) is found to decrease upon cycling. This occurs
primarily because of a reduction of sorbent reactive surface area
and porosity through sintering, either associated with the high
temperatures in the calcination environment or atomic rearrangement upon reaction to form CaCO3 and reform CaO. However, there
are other mechanisms, such as sulphation, ash fouling and mass
loss from the system, which can have a deleterious effect on sorbent performance. The loss in sorbent performance has successfully been modelled using a semi-empirical equation proposed by
Grasa et al. [4] (see Eq. (1)), where aN is the activity after N cycles
of carbonation (expressed in moles CO2 per mole of CaO) and calcination, kG can be seen as a decay constant and a1 can be seen as
the residual activity. Here, Eq. (1) will be used to project the carrying capacity past that tested.

aN

1
a1
1
1a

kG N
1

Several techniques have been proposed as methods of enhancing sorbent performance [1]. These generally take one of three
approaches: (i) to reduce the rate of decay of capacity as a result
of sintering, e.g., through doping [5], thermal pre-treatment [6]
or synthetic sorbents [7]; (ii) to increase the resistance to attrition,
e.g., through doping or synthetic sorbents; or (iii) periodically
increasing the sorbent reactivity, such as hydration [8,9]. In this
study, hydration is focussed upon.
Two potential methods of integrating reactivation through
hydration into a post-combustion CaL process are shown in
Fig. 1. The pressures and temperature in the carbonation and calci-

(a) CaL with reactivation by hydration


with direct carbonation
Hydrator
CaO+H2O Ca(OH)2
298 to 723 K, 101 kPa

Flue gas

Ca(OH)2

Carbonator
CaO+CO2 CaCO3
Ca(OH)2+CO2 CaCO3+H2O
923 K, 101 kPa

CaO

CaCO3

CaO

nation vessels are such that hydration is not favoured in these


reactors and an external hydration vessel is required; however,
in some higher pressure carbonation systems hydration can occur
in situ (typically pre-combustion systems, e.g., HyPr-RING [10]).
Hydration of calcined material can be performed using water,
water vapour or steam (see (Rn.2). The material from the hydrator
could be returned to either the carbonator or the calciner, resulting
in either direct carbonation of Ca(OH)2 (Rn.3) or indirect carbonation through a CaO intermediate (reverse of (Rn.2) followed
by (Rn.1)) respectively [11]. The hydration reactor has the potential
to operate at a high enough temperature to raise steam.
Early work on reactivation of CaO sorbent for CO2 was typically
performed using indirect carbonation [9,12,13]. The mechanism of
reactivation is typically described as occurring through an increase
in the reactive surface area [12], though indirect carbonation has
also been shown to result in crack formation and subsequently
the production of a very friable material [8,12,13]. Investigation
of direct carbonation has largely been inspired by a novel reactivation process proposed by Materic et al. [14] The mechanism for
particle dehydration has been under discussion; Materic et al.
[14] have proposed a superheated dehydration effect and Blamey
et al. [11] have proposed the formation of a CaCO3 shell around a
Ca(OH)2 core. Whatever the mechanism, direct carbonation of calcium hydroxide has been shown to result in higher conversions to
calcium carbonate and less friable material than from the use of
indirect carbonation [11,15].
Arias et al. [16] have modelled a CaL system for post-combustion CO2 capture using a scheme similar to that shown in
Fig. 1(a), where a portion of sorbent from the calciner was removed
and hydrated. Owing to the periodic hydration of portions of the
sorbent, particles could not be modelled using the typical empirical
equations used for modelling the decay in carrying capacity.
Instead, they modelled the decay of the carrying capacity of particles by using Eq. (1) modied to consider the carrying capacity in
the previous cycle and the effective number of cycles rather than
the actual number of cycles. They found that increasing the fraction of sorbent hydrated at each cycle could signicantly enhance
the average carrying capacity of sorbents in the system and subsequent capture efciency of CO2 from the carbonator. As a result,
the amount of material in the calcination and carbonation reactors
could be reduced if hydration were deployed.
Martinez et al. [17] have performed calculations for the balance
of moles steam generated per mole of CO2 captured for a series of

(b) CaL with reactivation by hydration


with in direct carbonation

O2 from air
separation unit

Flue gas

Coal

Calciner
CaCO3 CaO+CO2
1193 K, 101 kPa

Ca(OH)2

Carbonator
CaO+CO2 CaCO3
923 K, 101 kPa

Drying, compression,
liquefaction and storage

CaCO3

CaO

CaO

O2 from air
separation unit
Coal

Calciner
CaCO3 CaO+CO2
Ca(OH)2 CaO+H2O
1193 K, 101 kPa
CO2/H2O

CO2/H2O
N2, H2O and
other flue gases

Hydrator
CaO+H2O Ca(OH)2
298 to 723 K, 101 kPa

N2, H2O and


other flue gases

Drying, compression,
liquefaction and storage

Fig. 1. Simplied schematic of a CaL process for post-combustion CO2 capture with a fraction of the CaO from the calciner passed to a reactivator/hydrator with return of
Ca(OH)2 to (a) the carbonator and (b) the calciner, corresponding to (a) predominantly direct carbonation and (b) indirect carbonation; note that stoichiometric reactions are
shown for simplicity.

J. Blamey et al. / Fuel 150 (2015) 269277

experiments performed in a TGA investigating reactivation by


hydration. Their results show that often the largest improvements
in residual activity of sorbent come at the highest costs of steam
per mole of CO2 captured. This implies that there is a trade-off
between the improvement in performance of sorbent and the
objective of having a low steam consumption that will need to
be carefully analysed and optimised.
This paper follows up work performed on CaL with different
calcination temperatures in a uidised bed with reactivation by
ambient hydration (i.e., hydration in a humid atmosphere at room
temperature) and subsequent indirect carbonation of Ca(OH)2 [8].
Here, by comparison, CaL is investigated with different calcination
temperatures and steam hydration at elevated temperatures,
using a uidised bed. In addition, it investigates two different
modes of carbonation: direct and indirect carbonation of calcium
hydroxide. When it was noted that incomplete hydration
occurred in the case of the higher hydration temperatures, an
additional study of hydration behaviour of sorbents at different
hydration temperatures and extents of sintering was performed
using a TGA.
2. Experimental
Two sets of experiments were performed here
i. Fluidised bed hydration experiments. CaL cycles with three
different calcination temperatures were investigated. Sorbent was hydrated after 13 cycles using steam hydration
at two different temperatures, within the uidised bed. In
addition, both indirect and direct carbonation were
investigated.
ii. TGA hydration experiments. CaL cycles were performed in a
uidised bed, with the length of cycling varied to give a
wider range of pore volumes. Sorbent was removed from
the reactor and stored in a desiccator and steam hydration
was subsequently investigated in a TGA at three different
temperatures.

271

2.1. Fluidised bed hydration experiments


A diagram of the uidised bed (i.d. = 21 mm) apparatus used is
shown in Fig. 2. Limestone (in sand) was heated and subjected to
calcination and a further 13 cycles of carbonation and calcination
to generate spent sorbent. It was then cooled and hydrated in situ,
before being subjected to a further 5 cycles of carbonation and calcination. This resulted in a 1214 h continuous experiment, hence
13 cycles after hydration was not explored due to logistics, as had
been done in previous work [8].
Three sets of variables were investigated: (i) calcination temperatures before hydration: 1123, 1173 and 1223 K; (ii) hydration
temperatures: 473 and 673 K; and (iii) method of carbonation of
calcium hydroxide: direct and indirect.
2.1.1. Generation of spent sorbent
The generation of spent sorbent is described in detail elsewhere
[8]. Havelock limestone, 4 g of particle size fraction 500710 lm,
was mixed with 13 g of pre-heated sand at the calcination temperature. Sorbent was then cycled with a constant carbonation
temperature of 973 K and both calcination and carbonation times
of 900 s. The three calcination temperatures were chosen as 1123,
1173 and 1223 K, to allow comparison with prior work [8]. The
gas composition in both carbonation and calcination was 15% v/
v CO2, and balance N2, with a ow-rate of 47.5 cm3/s (which corresponds to a U/Umf  7 at 973 K using a correlation proposed by Wen
and Yu [18]). The gas ow-rates were set using calibrated rotameters. A non-dispersive infrared CO2 gas analyser (ADC SB100) was
used for online analysis of the CO2 concentration, which was
calibrated at the start of each experiment using pulses of pure N2
and calibration gas (15.00% v/v CO2, balance N2, BOC). Following
the nal calcination step, the hydration procedure was initiated.
2.1.2. Hydration procedure
Two hydration set-point temperatures were used, 473 and
673 K. The hydrator setup is shown in Fig. 2 and consisted of a syringe pump (Graseby 3100), which fed distilled water into 1/800 steel

Fig. 2. Fluidised bed reactor [8] modied for steam addition.

272

J. Blamey et al. / Fuel 150 (2015) 269277

tubing that was heated externally by controlled heating tapes.


Toward the end section of the hydrator, there was a three-way
valve such that the steam could either be vented to the atmosphere
or mixed with the inlet gas at the base of the reactor. The heating
tapes that pre-heated the steam and inlet gas were controlled to
523 K when in use. They were switched on 900 s prior to the calcination before hydration along with the syringe pump, set to 29.9
cm3/h and vented to atmosphere. At the end of the calcination,
the CO2 was switched off such that pure N2 owed into the reactor
(40.4 cm3/s at 293 K, 101 kPa). The power was switched off to the
uidised bed to allow cooling to hydration temperature. The analyser pump was switched off and gas-sampling line removed for the
duration of hydration to reduce the risk of water ingress to the
analyser. Approximately 300 s before the hydration set-point temperature was achieved, the syringe pump ow-rate was switched
to 99.9 cm3/h (still vented). This corresponds to a reaction owrate of steam of 37.6 cm3/s (equivalent ow-rate at 293 K,
101 kPa, assuming perfect gas). The ow-rate of steam was calculated by performing a mass balance on the water ow into and out
of the reactor, because the concentration of steam at the outlet was
not measured directly. Upon achieving temperature set-point, the
N2 ow-rate was switched to 16.9 cm3/s and the steam switched
to the reactor (to give 69% v/v steam). The sample was hydrated
for 900 s, after which an aliquot of sample was taken (using a
quartz rod with a quartz bucket attached, which would collect
3050 mg of sorbent/sand).
2.1.3. Further cycling
The procedure for further cycling depended on whether indirect
or direct carbonation of Ca(OH)2 was being investigated. In both
cases, following cycling, samples were removed under N2 and
placed in a desiccator to ensure full calcination and minimal hydration/recarbonation. The switching off and on of the analyser during
hydration changed the calibration slightly. A re-calibration of the
analyser for the cycles after hydration was therefore performed,
prior to removal of the sorbent, in a similar manner to that used
before the hydration experiment. The sorbent was removed under
nitrogen after the nal calcination and placed in a crucible in a desiccator to cool prior to mass measurement.
2.1.3.1. Indirect carbonation. Following hydration, the steam was
switched back to vent and the N2 ow-rate returned to 40.4 cm3/
s (the N2 ow-rate required for carbonation/calcination cycling).
The temperature was set to 1123 K for 900 s and the sorbent was
dehydrated. The CO2 was switched on 60 s before the end of the
dehydration period, when the temperature was set to 973 K and
5 carbonation/calcination cycles were performed (973 K for 900 s
and 1123 K for 900 s). At the end of the rst carbonation an aliquot
of sample was taken as a check on the CO2 uptake of the sorbent
because, following the switching back on of the analyser, the measured CO2 concentration took >300 s to stabilise fully.
2.1.3.2. Direct carbonation. This differed from indirect carbonation
in that the N2 ow-rate was returned to 40.4 cm3/s and CO2 was
switched on at the end of the hydration period (following sampling); only then was steam switched back to vent. Then, the temperature was set to 973 K and 5 carbonation/calcination cycles
performed under the same conditions as indirect carbonation, with
an aliquot of sample taken at the end of the rst carbonation.
2.1.4. Analysis of sample aliquots
The sample aliquots taken (i) at the end of the hydration period
and (ii) at the end of the rst carbonation period following hydration were subjected to analysis in a TGA (TA Instruments Q500) to
establish carbonation and hydration extents. In order to do this,
limestone was separated from the sand, manually, in a glove box

under an N2 atmosphere. To establish hydration and carbonation


extents, a three stage process was used: (1) heat to 383 K and hold
to remove any moisture from sample; (2) heat to 673 K and hold, at
this temperature the kinetics of dehydration are much faster than
those of decarbonation; and (3) heat to 1173 K and hold to achieve
full decarbonation of the sample. Using the masses at the end of
these stages (1) m0, (2) mh, (3) mf the conversion of the CaO
in the sample to CaCO3 and Ca(OH)2 were calculated (see Eqs. (2)
and (3)).

X CaOH2

X CaCO3

m0  mh
M H2 O

mh  mf
MCO2

mf F CaO
M CaO

mf F CaO
MCaO

2.1.5. Mass loss calculations


Mass measurements of the limestone added (m0), the sand
added (ms) and the nal calcined material at the end of the experiment (mf + ms) were taken to three decimal places. The mass loss,
expressed as a percentage, is calculated as shown in Eqs. (4) and
(5), where mt is the theoretical total mass of calcined limestone,
Mx is the molar mass of species x and FCaCO3 is the mass fraction
of CaCO3 in the original limestone as determined using X-ray
uorescence elemental analysis (Bruker AXS S4 Explorer). The
assumptions are (i) that there is no re-hydration or re-carbonation
of the calcined material and (ii) that the mass of sand remained
constant throughout the experiment.



m
Dmexp 100 1  f
mt
mt m0






F CaCO3 M CaO
1  F CaCO3
MCaCO3

2.2. TGA hydration experiments


Here, samples of Havelock limestone for hydration were prepared in the uidised bed (see Fig. 2). Hydration was then performed on the calcined sample ex situ in a TGA. Two sets of
variables were investigated: (i) number of cycles in the uidised
bed of 0, 2, 6 and 13; and (ii) hydration temperatures of 483, 578
and 678 K.
2.2.1. Cycling experiments
Sorbent was prepared for the TGA hydration experiments using
the method outlined in Section 2.1.1, with a calcination temperature of 1173 K. Samples were either cycled for 0 cycles (i.e.,
1 calcination with no carbonation), 2 cycles (1 calcination with 2
cycles of carbonation and calcination), 6 cycles or 13 cycles, before
being removed from the uidised bed following the nal calcination or placed in a desiccator. Then, the samples were sieved to
obtain sorbent particles >500 lm i.e., to separate the limestone
from the sand and stored in a vial within a desiccated jar prior
to steam hydration in the TGA.
2.2.2. Hydration experiments
Hydration was carried out in a TGA (Perkin Elmer TGA 7) on
18 0.1 mg aliquots of sorbent particles; which amounted to 12
layers of particles. Hydration was carried out at average temperatures of 483, 578 and 678 K. In each case, samples had hydrated slightly during transfer/separation, and, therefore, were rst
heated to 673 K at 0.83 K/s under N2, cold ow-rate 10 cm3/s, to
dehydrate before the temperature set-point was changed to the
hydration temperature. The furnace was turned off and the sample

J. Blamey et al. / Fuel 150 (2015) 269277

cooled prior to stabilisation at the desired temperature for 300 s.


Cooling and stabilisation took 300 s at 678 K, 1080 s at 578 K,
and 1620 s at 483 K. Then, steam, with an equivalent cold (293 K,
101 kPa) ow-rate of 32 cm3/s (76%, balance N2), was injected to
the system. The steam was generated using a syringe pump and
a steam generation system and was preheated before injection.
The steam system was preheated to 423 K and entered the TGA
at atmospheric pressure. Hydration was performed for a total of
1200 s. Repeats were performed for all TGA experiments performed at 483 and 678 K. Note that the partial pressures of steam
over Ca(OH)2 (see (Rn.2)) at equilibrium at 483, 578 and 678 K are
0.004, 0.32 and 7.8 kPa respectively [1,19]. The steam pressure of
reaction (77 kPa) is, therefore, in excess of the equilibrium
pressures.
2.3. Analytical techniques
Nitrogen adsorption (Micromeritics Tristar 3000 N2 Sorption
Analyser) analysis was performed to establish the BET surface area
and the BJH porosity associated with small pores.
3. Results and discussion
3.1. Fluidised bed experiments
3.1.1. Physical properties of the sorbent
The BET surface area and the BJH porosity associated with small
pores prior to hydration (i.e., after 13 cycles) are shown as a function of calcination temperature in Table 1. Data show that the BET
surface area and the porosity associated with small pores ([1 lm)

Table 1
Physical properties of cycled limestone, taken from nitrogen adsorption analysis;
conversion to Ca(OH)2 as a function of calcination temperature prior to hydration for
indirect and direct carbonation with hydration temperatures of 473 and 673 K.
Temperature
of calcination
(K)

BET
surface
area
(m2/g)

BJH
porosity [ 1 lm
(cm3/g)

Fractional
conversion of
CaO to
Ca(OH)2 at
473 K

Fractional
conversion of
CaO to
Ca(OH)2 at
673 K

1123
1173
1223

4.16
2.74
2.10

0.017
0.008
0.004

0.93
0.91
0.85

0.30
0.26
0.22

273

decrease with increasing calcination temperature. Note that the


surface area and porosities of the samples shown are for the calcined sorbent prior to hydration.
3.1.2. Carrying capacity from carbonation/calcination cycles
Data obtained for the carrying capacity of sorbents during the
experiments outlined are presented in Fig. 3 note that I denotes
indirect carbonation and D denotes direct carbonation, with
1123 K/1173 K/1223 K denoting the calcination temperature
before hydration. Data show that following hydration after 13
cycles there was a characteristic improvement in carrying capacity.
The sorbents showed higher carrying capacity following hydration
if (a) the calcination temperature before hydration was lower, (b)
the hydration temperature was lower, or (c) direct rather than
indirect carbonation was used. The carrying capacity remained
in excess of values predicted from application of the Grasa equation (Eq. (1) [4]) to data obtained prior to hydration for the ve
cycles following hydration tested; this occurred to variable extents
for all samples except those indirectly carbonated following hydration at 673 K.
The improvement of sorbent performance following hydration
is highlighted further in Fig. 4(a), which shows the increase in carrying capacity above the predicted value from the Grasa equation
(Eq. (1) [4]) when tted to the cycles prior to hydration. As a point
of reference, the reactivation extents from previous work with
ambient hydration [8] are also shown on the graph (I 293 K, indirect carbonation at room temperature). The reactivation extent is
essentially a composite of three factors: (i) the carrying capacity
without reactivation lower values of this will inate the reactivation extent; (ii) the carbonation extent (intrinsic carrying capacity)
of sorbent; (iii) the mass loss from the uidised bed, which will
decrease the observed carrying capacity articially in the cycle following hydration as it increases.
Data obtained for mass loss during the entirety of each experiment are shown in Fig. 4(b). Mass loss for the samples hydrated at
473 K was found to increase with increasing calcination temperature prior to hydration, in a similar manner to shown in previous work [8], but to a lesser extent. The samples that were
carbonated indirectly were found to exhibit more mass loss during
the course of an experiment than those directly carbonated. However, mass loss for the samples hydrated at 673 K appeared
relatively constant independent of calcination temperature before
hydration.

Fig. 3. Carrying capacity as a function of cycle number for hydration temperatures of (a) 473 K and (b) 673 K with calcination temperature varied before hydration (1123,
1173 and 1223 K) using indirect (I) and direct (D) carbonation as reactivation strategies also shown are ts of Eq. (1) to data obtained for calcination temperatures of 1123
and 1222 K before hydration.

274

J. Blamey et al. / Fuel 150 (2015) 269277

Fig. 4. (a) The carrying capacity increase following hydration in comparison to the predicted carrying capacity from tting Eq. (1) to data from cycling before hydration; (b)
mass loss as a function of calcination temperature throughout the experiment (including 13 cycles prior to hydration, hydration and 5 cycles post-hydration) including tting
from Eqs. (8) and (9) (dashed lines for direct and solid lines for indirect carbonation); Notation: indirect (I) and direct (D) carbonation with hydration temperatures of 473 and
673 K; reference data (I 293 K, indirect carbonation at room temperature) from Blamey et al. [8].

Table 2
Physical properties of cycled limestone, taken from nitrogen adsorption analysis.
Number of cycles

BET surface area (m2/g)

BJH porosity [ 1 lm (cm3/g)

0
2
6
13

16.17
8.77
5.15
2.88

0.118
0.093
0.056
0.027

3.1.3. Conversion as calculated from sample aliquots


The conversions to Ca(OH)2 at the end of the hydration period are
presented in Table 2. The trends observed are that the hydration
extent decreases upon increasing hydration temperature and
increasing calcination temperature prior to hydration for the
ranges of temperatures investigated the hydration temperature
shows greater inuence than the calcination temperature. The conversions to CaCO3 at the end of the carbonation period are presented
in Fig. 5(a). Data show that conversion increased upon decreasing
hydration temperature and for samples directly carbonated rather
than indirectly carbonated (D 473 K > I 473 K > D 673 K > I
673 K). Trends in carbonation extents as a function of calcination
temperature prior to hydration were harder to ascribe. It should
be noted that the conversion to CaCO3 obtained from the TGA

experiments on samples taken from the FBR and the conversion to


CaCO3 in the FBR normalised for mass loss at the end of the experiment (X CaCO3 exp=1  Dmexp ) showed good agreement.
3.1.4. Discussion
3.1.4.1. Trends in hydration extent. The hydration extent was found
to decrease substantially with increasing hydration temperature
and increase substantially with increasing porosity of material.
This was highlighted as an area for additional investigation and
is elaborated upon further in Section 3.2.
3.1.4.2. Trends in carbonation extent. A linear relationship was
found between the carbonation extent and the hydration extent
at the end of hydration, which was different for indirect and direct
carbonation techniques (see Fig. 5(b)). The slope was steeper for
direct carbonation, which achieves higher carbonation conversions
than indirect carbonation. Further investigation is needed to conrm a linear relationship because Fig. 5(b) is missing data from
0:4 < X CaOH2 < 0:8; however, it could be a useful correlation for
the purposes of modelling the reactivation process. The correlations are given in Eqs. (6) and (7) for indirect and direct carbonation respectively. The coefcients in these equations are
likely to be particular to the experimental set up, but the linear

Fig. 5. Conversion to CaCO3 following hydration (a) as a function of calcination temperature prior to hydration for indirect (I) and direct (D) carbonation with hydration
temperatures of 473 and 673 K; (b) as a function of hydration conversion to Ca(OH)2 upon hydration for direct and indirect carbonation.

275

J. Blamey et al. / Fuel 150 (2015) 269277

relationship may be general, as an analogous linear relationship


was found by Couturier et al. [20] for hydration extent prior to
sulphation.

X CaCO3 ;indirect 0:407X CaOH2 0:07

X CaCO3 ;direct 0:575X CaOH2 0:10

3.1.4.3. Trends in mass loss. Fig. 4(b) shows that mass loss during
experiments increased with increasing temperature of calcination
before hydration and with decreasing hydration temperature. It
also shows that indirect carbonation experiments had a higher
mass loss than direct carbonation experiments. It should be
emphasised that experiments performed here using steam hydration exhibited lower mass loss than those using ambient hydration
discussed previously [8]. Only two sets of experiments in the FBR
exhibited mass loss greater than 40%, which were for indirect carbonation at 473 K with calcination temperatures prior to hydration
of 1173 K and 1223 K.
The increase in mass loss with increasing calcination temperature prior to hydration can be explained using a similar argument to that developed previously [8]. Cycling with increased
calcination temperatures results in sorbent of increased density
and reduced porosity. This results in particles that have less
intra-particle space for Ca(OH)2 to grow into upon hydration. This
results in stress build-up upon hydration that is released after a
certain conversion, resulting in cracks developing in the particles.
At higher temperatures of hydration, it is possible that this stress
is released by annealing within the calcium hydroxide layer
formed, reducing the likelihood of fracture. If the sorbent is then
directly carbonated, carbonation will occur at a low temperature
(it is likely to start at the end of the hydration period, when the
gas stream is switched to CO2). Stress build-up in partially hydrated CaO particles has been observed by Molinder et al. [21], who
described nding stresses in the Ca(OH)2 phase more than 20 times
higher than its strength, leading to particle disintegration and generation of nano-sized crystallites. This will result in formation of
CaCO3, which has a larger molar volume than Ca(OH)2, that, with
the assistance of annealing/sintering processes at elevated temperatures, can grow across some of the cracks formed, giving the
particles structural rigidity. This process will be enhanced by further annealing upon heating, resulting in a sorbent that is not as
susceptible to fracture upon more vigorous uidisation at higher
temperatures. Indirect carbonation, however, goes through a CaO
intermediate, which will produce a denser, more porous, and more
fragile CaO skeleton that will be more susceptible to fracture upon
vigorous uidisation at higher temperatures, increasing mass loss.

The reduction in mass loss at higher hydration temperatures is


likely to be a result of the lower hydration conversions resulting
from pore blockage. This would result in less stress building up
in the particles. In addition, it is likely that the sintering processes
that enhance pore blockage will anneal the particles, reducing the
stress build-up. Empirical models for the reaction environment
have been developed to predict the mass loss as a function of surface area and hydration temperature (see Eqs. (8) and (9) for indirect and direct carbonation respectively), established through a
least squares minimisation.

Dmexp;indirect

Dmexp;direct

563
 0:047
T hyd SBET

215
0:134
T hyd SBET

3.2. TGA experiments


3.2.1. Physical properties of the sorbents
The BET surface area and the BJH porosity associated with small
pores are shown as a function of number of cycles in Table 2. Data
show that the BET surface area and the porosity associated with
small pores ([1 lm) decrease markedly upon cycling; the BJH porosity associated with small pores decreases more gradually than
the surface area. Note that the surface area and porosities of the
samples shown are for the calcined sorbent prior to hydration.
3.2.2. Hydration behaviour as a function of cycle number
Fig. 6 shows conversion to Ca(OH)2 of available CaO as a function of time for all samples tested (0, 2, 6 and 13 cycles) at
483 K, 578 K and 678 K under an atmosphere of 77 kPa steam. Note
that 100% conversion corresponds to 0.30 g H2O per g calcined sorbent. Data obtained at both temperatures show that it became progressively more difcult to hydrate the sample as the number of
cycles before hydration increased. At 483 K, the samples hydrated
after 0, 2, and 6 cycles were hydrated fully within 540 s; however, the sample hydrated after 13 cycles was not hydrated fully
within 1200 s. At 678 K, the samples hydrated after 0 and 2 cycles
hydrated very quickly; however, both the samples hydrated after 6
and 13 cycles appeared to reach a maximum/limiting conversion
after which the rate of reaction became very slow and the conversions to Ca(OH)2 were low after 1200 s. This effect can be seen as
pore blockage resulting in very slow gas-diffusion through pores
or solid-state diffusion. This is analogous to the slow phase of
the reaction discussed in relation to carbonation of CaO [5] and

Fig. 6. Conversion of CaO to Ca(OH)2 upon hydration as a function of time; comparing samples cycled to different extents at different temperatures.

276

J. Blamey et al. / Fuel 150 (2015) 269277

the effect referred to as the blocking effect described by Serris


et al. [22] This state is reached more rapidly for the samples cycled
to greater extents, because of their reduced porosity associated
with small pores (see BJH porosity of Table 2), which are more
prone to blocking upon formation of the less dense Ca(OH)2 (the
lower overall porosity will have an effect also see BJH porosity
data of Table 2). Experiments performed at 578 K exhibited intermediate behaviour.
3.2.3. Hydration conversion at 1140 s and maximum observed rate
Fig. 7 shows the conversion to Ca(OH)2 of available CaO sites at
1140 s, and maximum observed rate. The maximum conversion of
CaO to Ca(OH)2 at 1140 s tends to decrease with increasing cycle
number and increasing hydration temperature. The trends for the
maximum observed rate showed that the rate decreased with
increasing cycle number and decreasing hydration temperature.
It is interesting to note that increasing hydration temperature
results in an increased rate, but a reduced conversion to Ca(OH)2
after 1140 s. The exception is the rate for the sample calcined once
and hydrated at 678 K an effect that is not easily explained but
was reproducible.
3.2.4. Discussion
These data presented here have important consequences for the
application of hydration as a reactivation strategy to the eld of
CaL. Primarily: it cannot be taken for granted that the sample will
fully hydrate. This is important because incomplete hydration is
unlikely to produce a sorbent that is as reactive to CO2 as a sorbent
that fully hydrates. Improved hydration conversion was achieved
at lower hydration temperatures and for samples that had been
cycled fewer times i.e., that were less sintered.
3.2.4.1. Improved hydration conversion at lower hydration temperatures. A simplied schematic of CaL with a reactivation set
up is given in Fig. 1(a). A high hydration temperature will be
required, in order to allow the possibility of reclaiming higher
grade heat from the exothermic hydration reaction. However, this
work shows that higher hydration temperatures are detrimental to
the ability of a sorbent to hydrate. It is therefore, necessary that
careful process optimisation be performed to ensure that the
improvement in carrying capacity from performance of the sorbent
(and reduced cycling mass) upon hydration more than compensates for the increase in thermal load of sorbent and potential loss
of higher grade reclaimed.

3.2.4.2. Improved hydration conversion at lower cycle numbers. Here,


another balance is to be struck. At lower cycle numbers, higher
conversions to Ca(OH)2 and average rates of reaction can be
achieved and this is likely to correlate with subsequent reactivity
to CO2 (as is the case in the analogous reactivation for sulphation
process [20,23]). Higher conversions to Ca(OH)2 can be achieved
at lower cycle numbers, which is likely to result in high reactivity
sorbent; however, at lower cycle numbers, the sorbent is more
reactive to CO2 in any case and, therefore, less in need of reactivation. Again, the cycle number after which to hydrate will require
careful optimisation.
3.2.4.3. Projecting reactivation extent in a small uidised bed. The
maximum conversions to Ca(OH)2 have been used to predict the
subsequent conversions to CaCO3 in the reactor described in
Section 2.1. Here, only results for direct carbonation are shown.
The predicted conversion to CaCO3 of particles upon return to the
uidised bed for carbonation was calculated using Eq. (7). The
carrying capacity was then calculated to take into account material
lost from the bed using Eq. (9); results for predicted carrying capacity are shown in Fig. 8(a) with the predicted values from Eq. (1)
(i.e., without hydration). The carrying capacity increase across
hydration is shown in Fig. 8(b). This shows that, for a given
hydration temperature, the number of cycles can be optimised to
maximise reactivation extent, as indicated in Section 3.2.4.2, as
well as emphasising the negative effects of increasing the
hydration temperature.
3.2.4.4. Projecting to real systems. The empirical equations developed in this paper are particular to the experimental set-up used
(either the FBR or the TGA); however, we believe that the trends will
be indicative for other/larger systems. It should be noted that in real
systems, sorbent entering the hydrator from the calciner is likely to
be partially sulphated from sulphur impurities in the fuel in the calciner or ue gas entering the carbonator and potentially partially
carbonated as a result of the residence time distribution in the calciner. As a result, a less reactive sorbent to steam can be expected in
a real system; see Manovic et al. [24], for example. In addition, the
mass loss from the FBR is particular to the set-up, ow-rate and particle size chosen. Current pilot systems (see Arias et al. [25], for
example) under investigation are generally circulating uidised
beds with increased U/Umf and reduced particle size compared with
that investigated here. For these systems, the mass loss from the
system recorded here will be different; however, the indicated
reduction in particle size of increased mass loss in a small FBR will

(a) Conversion to Ca(OH)2 at 1140 s

(b) Maximum Rate

0.016

483 K
578 K
678 K

Maximum rate (dXCa(OH) /dt)

0.8

Maximum conversion to Ca(OH)2

0.6

0.4

483 K
578 K
678 K

0.2

10

Number of cycles before hydration

0.012

0.008

0.004

10

Number of cycles before hydration

Fig. 7. (a) Hydration conversion of CaO to Ca(OH)2 at 1140 s and (b) maximum observed rates of reaction (the differential of the conversion against time) as a function of cycle
number.

J. Blamey et al. / Fuel 150 (2015) 269277

277

Fig. 8. Predicted (a) conversions to CaCO3 against the predicted conversion from the Eq. (1) and (b) reactivation extent for materials hydrated in the section; assuming direct
carbonation and return to a small uidised bed, including mass loss.

have important ramications for the hydrodynamics of the process,


as well as the entrained nes lost from both the carbonator and the
calciner.
4. Conclusions
Steam hydration as a reactivation strategy for spent CaO-based
sorbents for CO2 capture has been investigated in both a uidised
bed reactor and a TGA. The uidised bed tests investigated steam
hydration at two different temperatures prior to either direct or
indirect carbonation. Direct carbonation was found to result in
material of a higher carrying capacity, because of increased carbonation conversion of individual particles and reduced mass loss.
However, increased hydration temperatures resulted in lower conversions to Ca(OH)2 and, subsequently, lower conversions to
CaCO3; though, the mass loss was reduced. Empirical equations
were used to provide relationships between (i) the conversion to
Ca(OH)2 and the subsequent conversion to CaCO3 and (ii) the mass
loss and the surface area prior to hydration and the prior surface
area and hydration temperature. These equations are particular
to the uidised bed used in this trial; however, the trends and ndings will be applicable to larger systems. Steam hydration of CaObased sorbent of various surface areas was then investigated in a
TGA, using three different hydration temperatures. This emphasised the relationship between hydration temperature and hydration conversions, as well as showing reduced hydration
conversions for increased cycle numbers. Projections from the
TGA data show that the cycle number can be optimised to maximise this relative increase in carrying capacity across hydration.
Acknowledgements
The authors wish to thank the Engineering and Physical
Sciences Research Council (UK) for funding the studentship of John
Blamey and CanmetENERGY for use of the thermogravimetric
analyser.
References
[1] Blamey J, Anthony EJ, Wang J, Fennell PS. The calcium looping cycle for largescale CO2 capture. Prog Energy Combust Sci 2010;36:26079.
[2] Snchez-Biezma A, Paniagua J, Diaz L, Lorenzo M, Alvarez J, Martnez D, Diego
ME, Abanades JC, et al. Testing postcombustion CO2 capture with CaO in a
1.7 MWt pilot facility. Energy Proc 2013;37:18.
[3] Strhle J, Junk M, Kremer J, Galloy A, Epple B. Carbonate looping experiments
in a 1 MWth pilot plant and model validation. Fuel 2014;127:1322.

[4] Grasa G, Abanades JC. CO2 capture capacity of CaO in long series of
carbonation/calcination cycles. Ind Eng Chem Res 2006;45:884651.
[5] Salvador C, Lu D, Anthony EJ, Abanades JC. Enhancement of CaO for CO2
capture in an FBC environment. Chem Eng J 2003;96:18795.
[6] Manovic V, Anthony EJ. Thermal activation of CaO-based sorbent and selfreactivation during CO2 capture looping cycles. Environ Sci Technol
2008;42:41704.
[7] Liu W, An H, Qin C, Yin J, Wang G, Feng B, et al. Performance enhancement of
calcium oxide sorbents for cyclic CO2 capture a review. Energy Fuel
2012;26:275167.
[8] Blamey J, Paterson NPM, Dugwell DR, Fennell PS. Mechanism of particle
breakage during reactivation of CaO-based sorbents for CO2 capture. Energy
Fuel 2010;24:460516.
[9] Manovic V, Anthony EJ. Steam reactivation of spent CaO-based sorbent for
multiple CO2 capture cycles. Environ Sci Technol 2007;41:14205.
[10] Lin SY, Suzuki Y, Hatano H, Harada M. Developing an innovative method, HyPrRING, to produce hydrogen from hydrocarbons. Energy Conver Manag
2002;43:128390.
[11] Blamey J, Lu DY, Fennell PS, Anthony EJ. Reactivation of CaO-based sorbents for
CO2 capture: mechanism for the carbonation of Ca(OH)2. Ind Eng Chem Res
2011;50:1032934.
[12] Fennell PS, Davidson JF, Dennis JS, Hayhurst AN. Regeneration of sintered
limestone sorbents for the sequestration of CO2 from combustion and other
systems. J Energy Inst 2007;80:1169.
[13] Hughes RW, Lu D, Anthony EJ, Wu YH. Improved long-term conversion of
limestone-derived sorbents for in situ capture of CO2 in a uidized bed
combustor. Ind Eng Chem Res 2004;43:552939.
[14] Materic V, Edwards S, Smedley SI, Holt R. Ca(OH)2 superheating as a lowattrition steam reactivation method for CaO in calcium looping applications.
Ind Eng Chem Res 2010;49:1242934.
[15] Materic V, Smedley SI. High temperature carbonation of Ca(OH)2. Ind Eng
Chem Res 2011;50:592732.
[16] Arias B, Grasa G, Abanades JC. Effect of sorbent hydration on the average
activity of CaO in a Ca-looping system. Chem Eng J 2010;163:32430.
[17] Martinez I, Grasa G, Murillo R, Arias B, Abanades JC. Evaluation of CO2 carrying
capacity of reactivated CaO by hydration. Energy Fuel 2011;25:1294301.
[18] Wen CY, Yu YH. A generalized method for predicting the minimum uidization
velocity. AIChE J 1966;12:6102.
[19] McBride BJ, Zehe MJ, Gordon S. NASA glenn coefcients for calculating
thermodynamic properties of individual species. Cleveland (Ohio,
US): National Aeronautics and Space Administration; 2002.
[20] Couturier MF, Marquis DL, Steward FR, Volmerange Y. Reactivation of
partially-sulfated limestone particles from a CFB combustor by hydration.
Can J Chem Eng 1994;72:917.
[21] Molinder R, Comyn TP, Hondow N, Parker JE, Dupont V. In situ X-ray diffraction
of CaO based CO2 sorbents. Energy Environ Sci 2012;5:895869.
[22] Serris E, Favergeon L, Pijolat M, Soustelle M, Nortier P, Gartner RS, Chopin T,
Habib Z. Study of the hydration of CaO powder by gassolid reaction. Cem
Concr Res 2011;41:107884.
[23] Shearer JA, Smith GW, Moulton DS, Smyk EB, Myles KM, Swift WM, Johnson I.
Hydration process for reactivating spent limestone and dolomite sorbents for
reuse in uidized-bed coal gasication. In: 6th International conference on
uidized bed combustion, Atlanta, Georgia, US; 1980. p. 101527.
[24] Manovic V, Lu D, Anthony EJ. Steam hydration of sorbents from a dual uidized
bed CO2 looping cycle reactor. Fuel 2008;87:334452.
[25] Arias B, Diego ME, Abanades JC, Lorenzo M, Diaz L, Martnez D, et al. Int J
Greenhouse Gas Control 2013;18:23745.

You might also like