You are on page 1of 25

Chemical Engineering Journal 283 (2016) 420444

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Review

High temperature CO2 sorbents and their application for hydrogen


production by sorption enhanced steam reforming process
Marziehossadat Shokrollahi Yancheshmeh, Hamid R. Radfarnia, Maria C. Iliuta
Department of Chemical Engineering, 1065 Av. de la Mdecine, Universit Laval, Qubec, Qubec G1V 0A6, Canada

h i g h l i g h t s
 SESR is a forefront technology to produce highly pure H2 in one step.
 High-temperature CO2 sorbents are the key element for successful SESR process.
 Ca-based and alkaline-based sorbents are the most investigated high-temperature CO2 sorbents.
 Capacity decay of Ca-based sorbents and slow kinetics of ceramics are the main challenges for industrial applications.
 Development of efcient hybrid catalyst/sorbent materials is an interesting opportunity for research.

a r t i c l e

i n f o

Article history:
Received 2 April 2015
Received in revised form 10 June 2015
Accepted 11 June 2015
Available online 10 July 2015
Keywords:
High-temperature CO2 sorbents
CaO-based materials
Alkaline-based materials
Hybrid catalystsorbent materials
Sorption enhanced steam reforming
Review

a b s t r a c t
Among the available techniques for hydrogen production, the sorption enhanced steam reforming (SESR)
is an emerging technology consisting in the integration of reforming reaction (H2 production) and selective separation (CO2 sorption) in a single step, to shift thermodynamically the reforming reaction and
increase hydrogen production. It is a forefront technology to produce highly pure hydrogen that has several advantages against the conventional steam reforming operation. The key element for a successful
SESR process is the selection of suitable high-temperature CO2 sorbents. Due to the weakness of current
CO2 sorbents (capacity decay and/or slow kinetics), the improvement of their performance is crucial to
make the SESR process interesting for industrial applications. This review focuses on the main characteristics and preparation methods of CaO-based and alkaline-based sorbents, their advantages and
drawbacks, the available techniques to improve their behavior in severe operating conditions, as well
as the progress of their application in two important SESR processes, namely sorption enhanced steam
methane reforming (SESMR) and sorption enhanced steam glycerol reforming (SESGR).
2015 Elsevier B.V. All rights reserved.

Contents
1.
2.

3.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
High temperature CO2 sorbents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
CaO-based sorbents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.1.
Application of various sources of calcium to produce CaO sorbent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.2.
Metal-stabilized CaO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.3.
Additional treatments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Ceramic CO2 sorbents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1.
Lithium zirconate (Li2ZrO3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.2.
Lithium orthosilicate (Li4SiO4) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.3.
Sodium zirconate (Na2ZrO3). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.4.
Other ceramic materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.5.
Kinetic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Hydrogen production by sorption enhanced steam methane reforming (SESMR) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Corresponding author. Tel.: +1 418 656 2204; fax: +1 418 656 5993.
E-mail address: maria-cornelia.iliuta@gch.ulaval.ca (M.C. Iliuta).
http://dx.doi.org/10.1016/j.cej.2015.06.060
1385-8947/ 2015 Elsevier B.V. All rights reserved.

421
421
421
423
424
427
431
431
433
434
434
435
435

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

4.
5.

3.1.
Application of CaO-based sorbents in SESMR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.
Application of ceramic sorbents in SESMR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Hydrogen production by sorption enhanced steam glycerol reforming (SESGR) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusion and recommendations for future works . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
Carbon dioxide (CO2) is a greenhouse gas (GHG) naturally present in the atmosphere. However, the increase of CO2 concentration due to human activities is the main factor that contributes
to Earths global warming. In May 2015 the CO2 concentration
reached 403 ppm, representing an increase by approximately 38%
in comparison with the middle of the 19th century [1]. Most
anthropogenic CO2 emissions come from the use of fossil fuels
for energy production in power plants, car engines and heating
processes. Hence, the continuous increase of global consumption
of energy is largely reected by the increase of CO2 emissions.
Among diverse possible approaches to lower the CO2 emissions,
the reduction of fossil fuel consumption by the increase of processes efciency and switch to less carbon-intensive fuels (e.g., natural gas) and/or to free-carbon energy sources (e.g., biomass) are
the most effective and economic methods to meet the goals of environmental rules [2]. In addition, the development of more efcient
and cost-competitive technologies to capture and storage the CO2
(CCS) has recently attracted considerable interest as an option to
control CO2 emissions. Post-combustion, pre-combustion and
oxy-combustion are the main methods for CO2 capture [3]. The
principle of post-combustion is to separate CO2 from combustion
exhaust gases that mainly contain CO2, N2 and O2. In the case of
oxy-combustion, pure oxygen is used rather than air for fuel combustion to avoid the dilution of CO2 by N2; as the resulting exhaust
practically contains only CO2 and water vapor, the CO2 can be easily
separated. The pre-combustion capture involves the gasication of
fossil fuels, which are rst converted into a mixture of hydrogen
and CO2; CO2 is therefore removed prior to combustion process.
Among several available technologies for CO2 capture like
chemical absorption, membranes and solid sorbents, the absorption based on chemical solvents (especially amine-based solutions)
is the most commonly used method for post-combustion CO2 capture due to its high CO2 removal efciency, particularly at low CO2
partial pressure. The use of solid sorbents is the most efcient technique for pre-combustion capture. Most available works in the
open literature are directed on high temperature solid sorbents like
lithium zirconate (Li2ZrO3) [4,5], sodium zirconate (Na2ZrO3) [6,7],
lithium silicate (Li4SiO4) [810] and CaO-based sorbents [1113].
High CO2 sorption capacity and adsorption/desorption kinetics
rate, relatively mild regeneration temperature, and multi-cycle stability are the most important parameters to be taken into account.
Therefore, there are still lots of challenges in the development of
efcient CO2 sorbents, especially related to long-term cyclic stability, capacity of sorption, regeneration condition and rate of CO2
capture.
Up to now, several research groups have provided review
papers about different strategies used to overcome the decay in
CO2 capture capacity of CaO-based sorbents over multiple carbonation/calcination cycles, including synthesis of CaO-based sorbents
from different precursors, stabilization of sorbents by incorporating CaO in support materials, and reactivation of sorbents with
hydration, thermal pretreatment and chemical pretreatment [14
17]. Wang et al. [18] provided a review on the synthesis methods,
CO2 adsorption/desorption characteristics and possible sorption
mechanisms for LixZryOz with different Li/Zr ratios, but it did not

421

436
438
439
440
441
441

discuss other types of ceramic CO2 sorbents such as Na2ZrO3 and


Li4SiO4. Although the sorption enhanced steam reforming is one
of the most important applications of high-temperature sorbents,
no comprehensive review is available in the open literature. This
review paper focuses on the main characteristics and preparation
methods of all appropriate high-temperature CO2 sorbents, the
available techniques for improvement of their behavior in severe
operating conditions and the progress of their application in the
SESR process. It is divided into the following main sections: (i)
CaO-based sorbent; (ii) ceramic CO2 sorbents; (iii) sorption
enhanced steam methane reforming (SESMR); and (iv) sorption
enhanced steam glycerol reforming (SESGR).
2. High temperature CO2 sorbents
A highly efcient sorbent for CO2 capture at high temperature
should possess specic properties such as: thermal stability at high
operating temperatures (450700 C); adequate CO2 sorption
capacity and kinetics; easiness of sorbent regeneration;
long-term cyclic stability; and reasonable production cost. The
most promising high-temperature solid sorbents available in the
literature mainly include ceramic alkaline-based and CaO-based
sorbents. Hydrotalcite (HTLc) is another type of sorbent, but its
CO2 uptake capacity is very low in comparison to the other materials [19,20]. The following section will therefore review the properties and preparation methods of the most commonly used solid
sorbents, including CaO-based and alkaline-based sorbents.
2.1. CaO-based sorbents
One cannot deny that CaO is the most famous natural CO2 sorbent that exists in nature in the forms of limestone (CaCO3) and
dolomite (CaMg(CO3)2). This sorbent has attracted a lot of attention because of its low raw material cost, high CO2 sorption capacity and adequate kinetics of reactions. The sorption/desorption
reaction of CaO is given by:

CaOs CO2 g $ CaCO3 s DH 298 175:7 kJ=mol

Theoretical (stoichiometric) CO2 capture capacity of CaO is as


high as 0.786 g of CO2/g of sorbent. Nevertheless, dolomite
(CaMg(CO3)2) and huntite (CaMg3(CO3)4) have lower CO2 sorption
capacity (dolomite: 0.46 g of CO2/g of sorbent and huntite: 0.25 g
of CO2/g of sorbent) because MgO does not participate in CO2
adsorption.
The endothermic regeneration reaction (reverse reaction in Eq.
(1)) needs a high amount of energy and usually occurs at a temperature above 900 C in CO2 atmosphere [21]. Although CaO has
some benets as a CO2 sorbent, its industrial application encounters with some critical issues such as the loss of sorption capacity
in long-term operation and the loss of reactivity with sulfur containing gases to form CaSO4 [11,2224]. The loss of CO2 sorption
capacity during cyclic operation is mainly resulted from the sintering phenomenon, which consists of the agglomeration of small
particles, the change of pore shapes, and the pore shrinkage. As it
can be seen in Fig. 1, the amount of unreacted CaO increases along
with the cycle number until a rigid interconnected CaO skeleton is

422

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

Fig. 1. Sintering phenomenon of CaO; light grey: CaO phase and dark grey: CaCO3 phase [26].

Fig. 2. Typical weight changes vs. time for a repeated number of calcination/carbonation cycles of Piaseck limestone [23].

formed after 50 cycles. The carbonation takes place therefore on


the external surface of the material. Reactive dynamics simulations
with the reactive force eld (ReaxFF) performed in NVE ensembles
showed that the sintering of CaO particles has occured because of
their expansion during CO2 adsorption [25]. This expansion, which
leads to rapid sintering, is strongly inuenced by temperature and
particle separation distance: the higher the adsorption temperature and the shorter the distance between two sorbent particles,
the faster the sintering rate during the adsorption process.
Therefore, the capacity decay of CaO sorbents during multiple carbonation/calcination cycles depends on the experimental temperature, the precursor type, and the duration of the recarbonation
step [26].
The CO2 chemisorption on CaO generally consists of two
regimes, fast and slow kinetic steps. Dou et al. [27], Rout et al.
[28] and Mohammadi et al. [29] performed kinetic studies on the
carbonation reaction of different Ca-based synthetic sorbents and
found that the carbonation reaction was controlled by both chemical reaction at the CaOACaCO3 interface and carbonate layer diffusion. The fast kinetic step, which is controlled by chemical reaction,
will be continued untill the carbonate layer surrounding the unreacted CaO core is completed. Then, the slow kinetic step that is
controlled by gas diffusion starts. At this step, the product layer
restricts the access of CO2 molecules to reactive sites. Alvarez
and Abanades [30] reported that the gas diffusion and thus, the
sorption rate, can be limited above a critical carbonate layer thickness of 50 nm. Mostafavi et al. [31] found that the CaO sorbent
derived from limestone represented a higher initial carbonation
reaction rate in comparison with that derived from dolomite,
because of excessive CaO active sites. On the other hand, the ultimate conversion was higher for dolomite derived CaO sorbent at

low temperature (550 C), whereas this value was higher for limestone derived CaO sorbent at high temperatures (600675 C). As a
matter of fact, at the beginning of the reaction, the growth of CaCO3
layer for limestone was faster than that of dolomite because of the
higher rate of reaction. Since Knudsen diffusion is related to the
square root of temperature, the diffusion through CaCO3 layer
enhanced as the temperature increased above 550 C. Therefore,
limestone showed a higher ultimate conversion at 600 and 675 C.
Some researchers studied the CO2 sorption behavior of limestone in long-term cyclic operation. Grasa and Abanades [23] evaluated the CO2 capture performance of limestone over 500
carbonation/calcination cycles and observed that the CO2 uptake
capacity signicantly decreased during the rst 20 cycles and then
stabilized along with the cycle number around 0.0750.08 residual
conversion up to 500 cycles. For CaO sorbent, Fig. 2 shows the
decrease of CO2 capture capacity along with the cycle number, as
well as the kinetic steps. In another study, Sun et al. [24] examined
the CO2 sorption behavior of limestone through more than 1000
carbonation/calcination cycles. They reported a calcium conversion
between 4% and 17% (depending on the carbonation time) after
150 cycles.
Some kinds of natural CaO sorbent containing MgO, such as
dolomite and huntite, possess better stability during cyclic operations. Silaban et al. [11] observed that after six adsorption/desorption cycles, the capture capacity of limestone was decreased from
61% to 35% of its theoretical value. However, dolomite was already
stable at a sorption capacity of 40%. The behavior of dolomite was
ascribed to the presence of MgO in its structure, which provided a
better structural stability. Bandi et al. [32] examined the CO2 capture activity of calcite, dolomite and huntite for 47 adsorption/desorption cycles. The experimental results showed that huntite could

423

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444


Table 1
Summary of investigations on CaO sorbents synthesized from different precursors.
Precursor

Reaction conditions
Ads.

Reactor

Number
of cycles

TGA
TGA
TGA
TGA
TGA
TGA
TGA

27
17
9
57
100
60
20

0.49
0.5
0.66
0.19
0.17
0.39
0.39

[13]
[13]
[39]
[39]
[30]
[45]
[45]

FB
FB
TGA
TGA
TGA
TGA
TGA
TGA

9
9
11
20
20
9
9
9

0.303
0.285
0.40
0.51
0.2
0.32
0.1
0.25

[40]
[40]
[34]
[46]
[46]
[35]
[35]
[35]

Reg.

Calcium acetate
Calcium acetate
Calcium D-gluconate
Calcium D-gluconate
Nano-CaCO3 (40 nm)
Calcium naphthenate (FSP-made CaO)
Calcium naphthenate (FSP-made CaO)

700 C,
700 C,
650 C,
650 C,
650 C,
700 C,
700 C,

30%
30%
15%
15%
15%
30%
30%

CO2,
CO2,
CO2,
CO2,
CO2,
CO2,
CO2,

CaAc2
CaCO3
CaCO3
Ca(NO3)24H2O
Ca(NO3)24H2O
CaO
CaCO3
Ca(OH)2

650 C,
650 C,
650 C,
650 C,
650 C,
750 C,
650 C,
650 C,

15.30% CO2
15.30% CO2
15% CO2, 2 h
15% CO2, 15 min
15% CO2, 15 min
100% CO2, 40 min
100% CO2, 30 min
100% CO2, 40 min

300 min
10% H2O, 300 min
30 min
30 min
20 min
300 min
5 min

preserve around 84% of its initial capacity, while dolomite and calcite were able to maintain 55% and 38% of their initial capacity,
respectively. However, huntite possessed less CO2 uptake capacity
in comparison to dolomite and calcite because of its higher content
of inactive MgO. According to these studies, the deactivation of
CaO sorbent derived from natural sources is unavoidable.
Different strategies have therefore been proposed to improve the
CO2 sorption performance of CaO sorbent, such as (i) application
of various sources of calcium to produce CaO sorbent, (ii) incorporation of stable inert materials into CaO structure, and (iii) reactivation and treatment of sorbent.
2.1.1. Application of various sources of calcium to produce CaO sorbent
Many research groups have focused on the production of CaO
sorbents from various calcium precursors, including CaCO3 [33
35], Ca(OH)2 [3537], organometallic [13,3843] and nano-sized
CaO/CaCO3 [39,42,4449], with the aim of providing a CaO-based
sorbent with meso- and macroporous structure, high specic surface area, large pore volume, and small particle size (Table 1).
These specications are mandatory for a sorbent with high and
stable CO2 capture capacity. The performance of CaO sorbents
derived from two most promising groups, organometallic and
nano-sized CaO/CaCO3 precursors are further discussed.
2.1.1.1. Organometallic precursors. Up to now, a number of studies
have reported the production of CaO sorbents from various
organometallic precursors (OMPs) and the relation between the
structural characteristics of OMPs-derived CaO sorbents with their
CO2 uptake performance [13,3843]. The decomposition of OMPs
leads to the formation of meso- and macroporous structures with
large surface area, which results in the enhancement of CO2 capture performance [13,39]. In an earlier work, Silaban et al. [43]
found that CaO synthesized from calcium acetate possessed a
higher CO2 capture capacity than CaO synthesized from calcium
carbonate independent of the calcination temperature. Similarly,
Lu et al. [13] examined the CO2 capture activity of the CaO sorbents
derived from calcium acetate monohydrate, calcium carbonate,
calcium hydroxide, and calcium nitrate tetrahydrate. According
to the experimental results, the CaO sorbent obtained from calcium
acetate demonstrated the best performance with a CO2 sorption
capacity of 0.49 g of CO2/g of sorbent after 27 cycles (carbonation
at 700 C under 30% CO2/He balance and calcination at 700 C
under He), because of its large BET surface area and pore volume.
The SEM images revealed that this sorbent contained a uffy structure, which contributed to its high surface area and large pore volume. In a further study, Liu et al. [39] prepared CaO sorbents using

700 C, 100% He, 30 min


700 C, 100% He, 30 min
900 C, 100% N2, 10 min
920 C, 15% CO2, 2 min
850 C, 100% N2, 10 min
700 C, 100% He, 30 min
nonisothermal to 900 C,
100% He, 40 min
850 C, 100% N2
850 C, 100% N2
950 C, 100% N2
850 C, 100% N2, 10 min
950 C, 100% CO2, 10 min
750 C, 100% N2, 30 min
750 C, 100% N2, 30 min
750 C, 100% N2, 40 min

CO2 uptake at
last cycle (g-CO2/g-ads)

Ref.

different precursors, including calcium acetate hydrate, calcium


citrate tetrahydrate, calcium D-gluconate monohydrate, calcium
formate, calcium L-lactate hydrate, calcium hydroxide, microsize
calcium carbonate, nanosize (<70 nm) calcium carbonate, and
nanosize (<160 nm) calcium oxide. Among the developed materials, the CaO sorbent derived from calcium D-gluconate monohydrate exhibited the highest CO2 capture capacity of 0.66 g of
CO2/g of sorbent at the 9th cycle. Yang et al. [40] synthesized four
types of Ca-based sorbents from calcium acetate monohydrate, calcium carbonate, calcium hydroxide, and calcium oxide precursors
by calcination and hydration reactions. The cyclic CO2 capture
experiments showed that CaO sorbents derived from calcium acetate and calcium carbonate presented a higher CO2 sorption capacity (0.299 and 0.284 g of CO2/g of sorbent at 650 C, respectively)
compared to CaO sorbents derived from calcium oxide and calcium
hydroxide due to the larger pore volume and higher specic surface area of the former ones. According to the cyclic carbonation/decarbonation experiments, the CO2 sorption capacity of
CaAc2ACaO and CaCO3ACaO sorbents increased in the rst cycle
and then decreased and reached to 0.303 and 0.285 g of CO2/g of
sorbent after 9 cycles. Grasa et al. [38] studied the cyclic CO2 sorption performance of CaO sorbents prepared from calcium hydroxide, calcium acetate, and calcium oxalate under realistic
calcination conditions (regeneration at temperatures around
900 C under CO2 atmosphere). Although these synthetic sorbents
performed well under mild reaction conditions, they showed a dramatic decay in CO2 capture capacity under realistic severe calcination conditions, the behavior being similar to natural limestone.
The best synthetic sorbent, which was derived from calcium acetate, exhibited a nal CO2 sorption uptake slightly higher than that
of limestone. This CO2 capture behavior did not justify the higher
cost of sorbent production from chemical precursors.

2.1.1.2. Nano-sized CaO and CaCO3. For porous CaO with critical particle size less than 44 nm or a large single CaO crystal with critical
crystal size less than 220 nm, the carbonation reaction completes
within the fast kinetically-controlled regime and the slow
solid-state diffusion controlled regime does not exist [47,50].
Therefore, different research groups studied nano-sized CaO and
CaCO3 as calcium precursors [39,42,4449]. All of these studies
show that nano-sized sorbents possess much better CO2 capture
activity in comparison to micro-sized CaO. However, there are
two primary hurdles for using the nano-sized CaO sorbents in practical applications: (i) the nano-sized CaO sorbents synthesized
from nano-sized particles are susceptive to sintering because of
higher surface area [44]; (ii) the methods used for the synthesis

424

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

Table 2
Summary of investigations on metal oxide stabilized CaO-based sorbents.
Stabilizer

Stab.
content (wt%)

Reaction conditions

Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca9Al6O18
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Ca12Al14O33
Al2O3
Al2O3
Al2O3
Ca9Al6O18
Ca9Al6O18
Ca9Al6O18
Ca9Al6O18
Ca9Al6O18
Ca9Al6O18
Ca3Al2O6
Ca3Al2O6
Cement
MgO
MgO
MgO
MgO
MgO
MgO
CaZrO3
CaZrO3
CaZrO3
CaZrO3
CaZrO3
CaZrO3
CaZrO3
CaZrO3
TiO2
TiO2
Mesoporous silica shell
Silica
Y2O3
Y2O3

25
25
25
25
25
65
15
7.7
7.7
9.2
20
42
42
7.7
25
14.7

?
39
19
19
8
20
20
20
7.5
7.5
22.1
34
9
10
26
25
25
25
25
25
58
10
30
30
N.a.
30
26.2
58.16
10

43
14.7
20
20

690 C,
690 C,
700 C,
700 C,
690 C,
650 C,
750 C,
690 C,
650 C,
750 C,
650 C,
700 C,
850 C,
650 C,
700 C,
675 C,
550 C,
600 C,
600 C,
750 C,
700 C,
710 C,
650 C,
650 C,
650 C,
650 C,
650 C,
650 C,
690 C,
650 C,
850 C,
758 C,
700 C,
750 C,
650 C,
650 C,
700 C,
700 C,
650 C,
650 C,
650 C,
800 C,
700 C,
650 C,
600 C,
600 C,
400 C,
675 C,
675 C,
650 C,
650 C,

Ads.

Reactor

Number
of cycles

CO2 uptake at last


cycle (g-CO2/g-ads)

Ref.

TGA
TGA
TGA
TGA
TGA
TGA
FB
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
PB (Packed-bed)
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
PB (packed bed)
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA

13
13
50
56
45
50
20
30
30
30
28
100
100
30
100
31
10
150
40
30
10
15
35
35
50
31
31
25
45
100
30
50
60
1250
24
44
50
100
30
30
20
90
1050
25
15
40
1
31
31
10
10

0.45
0.333
0.41
0.22
0.26
0.19
0.26
0.37
0.22
0.55
0.51
0.4
0.25
0.6
0.4
0.27
0.34
0.130.15
0.4
0.36
0.31
0.62
0.52
0.2
0.48
0.57
0.33
0.33
0.45
0.35
0.17
0.53
0.44
0.17
0.56
0.46
0.54
0.3
0.37
0.31
0.48
0.34
0.30
0.29
0.15
0.23
0.2
0.24
0.15
0.57
0.49

[12]
[12]
[65]
[65]
[66]
[67]
[68]
[76]
[80]
[79]
[63]
[69]
[69]
[52]
[74]
[61]
[194]
[72]
[83]
[75]
[75]
[73]
[195]
[195]
[51]
[78]
[78]
[71]
[77]
[64]
[125]
[86]
[53]
[54]
[87]
[55]
[88]
[45]
[92]
[92]
[29]
[93]
[91]
[71]
[90]
[56]
[58]
[61]
[61]
[62]
[62]

Reg.
14% CO2, 30 min
14% CO2, 30 min
20% CO2, 30 min
20% CO2, 30 min
15% CO2, 30 min
33% CO2, 10 min
14% CO2, 8.3 min
15% CO2, 30 min
15% CO2, 10 min
40% CO2,20 min
15% CO2, 30 min
30% CO2, 10 min
100% CO2, 10 min
15% CO2, 30 min
90% CO2
100% CO2, 20 min
7% CO2, 27% H2O, 12 min
20% CO2, 25 min
100% CO2, 10 min
40% CO2, 20 min
40% CO2, 20 min
100% CO2, 30 min
15% CO2, 30 min
15% CO2, 30 min
0.015 MPa CO2, 30 min
100% CO2, 30 min
100% CO2, 30 min
15% CO2, 30 min
15% CO2, 30 min
20% CO2, 30 min
100% CO2, 10 min
100% CO2, 30 min
90% CO2
25% CO2, 20 min
15% CO2, 30 min
15% CO2, 30 min
20% CO2, 10 min
30% CO2, 30 min
15% CO2
100% CO2
100% CO2, 60 min
50% CO2, 5 min
100% CO2, 30 min
15% CO2, 30 min
100% CO2, 30 min
20% CO2, 10 min
100% CO2, 30 min
100% CO2, 20 min
100% CO2, 20 min
20% CO2, 30 min
20% CO2, 30 min

850 C, 100% N2, 10 min


950 C, 20% CO2, 10 min
850 C, 100% N2, 5 min
980 C, 100% CO2, 5 min
850 C, 100% N2, 10 min
800 C, 100% N2
750 C, 100% N2, 8.3 min
850 C, 100% N2, 5 min
900 C, 15% CO2, 5 min
750 C, 100% N2, 20 min
800 C, 100% N2, 10 min
700 C, 100% He, 10 min
950 C, 30% CO2, 10 min
900 C, 100% N2, 10 min
700 C, 100% He
850 C, 100% N2, 10 min
750 C, 100% N2, 15 min
1000 C, 86% CO2, 15 min
700 C, 100% N2, 8 min
750 C, 100% N2, 20 min
925 C, 100% CO2, 20 min
950 C, 100% N2, 5 min
800 C, 100% N2, 10 min
1000 C, 80% CO2, 10 min
800 C, 100% N2, 10 min
800 C, 100% N2, 10 min
930 C, 100% CO2, 5 min
750 C, 100% Ar, 30 min
800 C, 100% N2, 5 min
850 C, 100% N2, 10 min
850 C, 100% N2, 10 min
758 C, 100% He, 30 min
700 C, 100% He
750 C, 100% N2, 30 min
900 C, 100% N2, 10 min
900 C, 100% N2, 10 min
730 C, 100% N2, 10 min
700 C, 100% He, 30 min
800 C, 100% air
950 C, 100% CO2
700 C, 100% N2, 20 min
800 C, 100% N2, 15 min
700 C, 100% He, 30 min
750 C, 100% Ar, 30 min
750 C, 100% Ar, 30 min
750 C, 100% N2, 10 min
400 C, 100% N2
850 C, 100% N2, 10 min
850 C, 100% N2, 10 min
850 C, 100% N2, 5 min
950 C, 100% CO2, 5 min

of nanoparticles, such as ame spray pyrolysis and solgel method,


are relatively complex and costly. One of the best nano-sized sorbents was developed by Luo et al. [46] from calcium nitrate
tetrahydrate and citric acid monohydrate precursors by the sol
gel method. They observed well-dispersed uniform particles
(200 nm) within this new sorbent. This sorbent showed the CO2
adsorption capacity of 0.51 g of CO2/g of sorbent under mild calcination conditions and 0.20 g of CO2/g of sorbent under severe calcination conditions over 20 cycles, which are signicantly higher
than those for CaO sorbents synthesized from commercial microand nano-sized CaCO3. In addition, this new sorbent exhibited very
high reaction rate during the carbonation (60% calcium conversion
ratio within 20 s) as well as a better sintering-resistant property.

sorbents. As a result, many research groups have focused to nd


a new technique to improve the stability of Ca-based sorbents.
The incorporation of inert support materials, including aluminum
oxide (Al2O3) [51,52], magnesium oxide (MgO) [5355], zirconium
oxide (ZrO2) [29], titanium oxide (TiO2) [5658], silica (SiO2) [59
61], yttrium oxide (Y2O3) [62], etc., into the sorbent structure is an
efcient technique to improve the stability of CaO-based sorbents
derived from synthetic precursors (Table 2). The support material,
which possesses a high Tammann temperature, is dispersed among
the CaO particles during the synthesis and inhibits the CaO grain
sintering during carbonation/calcination cycles. In the following
sections, we provide an overview on the studies concerning the
most promising supporting materials: Al2O3, MgO and ZrO2.

2.1.2. Metal-stabilized CaO


As mentioned above, the loss of CO2 capture capacity during
cyclic carbonation/calcination operations is a main problem even
when the synthetic precursors are used to prepare CaO-based

2.1.2.1. Al-stabilized CaO. Al-stabilized CaO sorbents are the most


studied
metal-stabilized
CaO-based
sorbents.
Different
Al-stabilizer phases can be produced (Al2O3, Ca12Al14O33,
Ca9Al6O18, and Ca3Al2O6), depending on calcium and aluminum

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

precursors as well as synthesis method [12,63,64]. The inert support materials are distributed between the CaO particles during
the synthesis procedure and prevent the sintering phenomenon
during multiple carbonation/calcination cycles resulting in the
enhancement of CO2 capture performance. One of the earliest
reports about the synthesis of Al-stabilized CaO-based sorbents
has been published by Li et al. [12,65]. They developed
Ca12Al14O33-stabilized CaO sorbent from aluminum nitrate nonahydrate (Al(NO3)39H2O) and calcium oxide precursors via wet
mixing method. The synthesized sorbent consisting of 75 wt.%
CaO and 25 wt.% Ca12Al14O33 presented the highest adsorption
capacity of 0.41 g of CO2/g of sorbent at the end of the 50th cycle
under mild calcination conditions (850 C, 100% N2). Using more
realistic calcination conditions (980 C, 100% CO2) reduced the
adsorption capacity to 0.22 g of CO2/g of sorbent after 56 cycles.
However, it is still more than the CO2 capture capacity of dolomite
(0.16 g of CO2/g of sorbent).
Encouraged by the promising results reported by Li et al.
[12,65], several attempts have been made to incorporate aluminum compounds to the CaO structure, mainly focusing on the
application of different calcium and aluminum precursors
[63,66,67], optimization of Ca/Al ratio [6870], and use of different
synthesis methods [64,7180].
Zhou et al. [63] employed a wet-mixing technique to synthesize
Al-stabilized CaO sorbents from various calcium and aluminum
precursors. They found that different inert support materials,
including Al2O3, Ca12Al14O33 or Ca9Al6O18, could be produced
depending on calcium and aluminum precursors used during the
synthesis process. According to the experimental results, most of
the Al-stabilized CaO sorbents showed higher CO2 sorption capability and stability during multi-cyclic carbonation/calcination
operation in comparison to pure CaO, which was attributed to
the bimodal pore size distribution with an adequate number of
small pores, the high specic surface area of sorbents, and the uniform distribution of inert support materials among CaO particles.
Among the developed sorbents, CaO/Ca9Al6O18 derived from calcium citrate and aluminum nitrate presented the best performance
with the CO2 capture capacity of 0.51 g of CO2/g of sorbent after 28
cycles. The authors also proposed a formation mechanism for the
support materials (Fig. 3). The transition between steps 3 and 4
depends on the precursors used during the synthesis step, which
can limit Ca2+ diffusion into the stabilizer structure for further
reaction.
Pacciani et al. [68,70] prepared a number of CaO/Ca12Al14O33
synthetic sorbents with different CaO/inert material ratios.
Among the studied materials, the sorbent consisting of 85 wt.%
CaO and 15 wt.% Ca12Al14O33 showed the highest activity in a uidized bed reactor with a CO2 capture capacity of 0.26 g of CO2/g
of sorbent after 20 cycles (adsorption under 14% CO2 ow and
regeneration under pure N2), which decreased to 0.17 g of CO2/g
of sorbent after 110 cycles. In another study, Koirala et al. [69]
applied the single nozzle ame spray pyrolysis method to synthesize Al-stabilized CaO sorbents with different Al/Ca ratios. They
found that a higher Al doping into CaO improved the sorbent stability. According to the multiple carbonation/calcination experiments, the synthetic sorbent with the Al/Ca molar ratio of 3:10
exhibited the best performance with a CO2 uptake capacity of
0.40 g of CO2/g of sorbent, which was stable over 100 cycles. The
improved durability of Al-stabilized CaO sorbents was owing to
the uniform distribution of Ca12Al14O33, which was produced in

(1)
(2)
Ca & Al precursors
CaCO3 +Al2 O3
CaO+Al2 O3
CO2 , H 2O
CO2
( 3)
( 4)

Ca12 Al14 O33


Ca 9 Al6 O18
CaO

Fig. 3. Possible mechanism for Al-stabilizer formation (adapted from [63]).

425

the synthesis step, among the CaO particles. The sorbent with the
Al/Ca molar ratio of 3:10 also performed well under severe calcination conditions, indicating a residual sorption capacity of 0.25 g
of CO2/g of sorbent at the 100th cycle. These results revealed that
the presence of CO2 in the regeneration step led to the accelerated
structural sintering.
Recently, different research groups have developed various synthesis methods with the aim of increasing surface area and pore
volume and obtaining high dispersion of inert supports within
the sorbent structure: wet mixing [12,63,65,66,68,70], limestone
acidication by citric acid followed by two step calcination [71],
solid-state reaction [72], ultrasonic spray pyrolysis (USP) [73],
combination of precipitation and hydration [74], co-precipitation
[75], citrate preparation [64], solgel [77,79], citrate-assisted sol
gel technique followed by two-step calcination [78], single nozzle
ame spray pyrolysis (FSP) [69], and precipitation [80]. Some of
these techniques result in Al-stabilized CaO sorbents with a great
performance in cyclic carbonation/calcination experiments.
Sayyah et al. [73] developed a series of Al-stabilized CaO sorbents
using ultrasonic spray pyrolysis (USP) method. The precursors, calcium nitrate tetrahydrate (Ca(NO3)24H2O) and aluminum nitrate
nonahydrate (Al(NO3)39H2O), were solved in ethanol. The
obtained solution was nebulized via ultrasound, followed by carrying through a furnace tube (set at 600 C) by argon gas. The synthesized sorbents were tested in carbonation (710 C, 100% CO2) and
calcination (950 C, 100% N2) cyclic operations. The results showed
that the optimum value for Al/Ca ratio is around 0.08. The CO2
adsorption capacity of this sorbent decreased from 0.65 g of
CO2/g of sorbent to 0.62 g of CO2/g of sorbent over 15 calcination/carbonation cycles. The high performance of the sorbents showed
the capability of USP method in homogenous dispersion of additives in the CaO matrix and production of materials with high surface area and high stability. Zhang et al. [64] applied the citrate
preparation route to synthesize a Ca-based sorbent consisting of
9 wt.% Al2O3 and 91 wt.% CaO from aluminum nitrate, citric acid,
and CaCO3. The developed sorbent was activated by a four-step
activation procedure instead of the common one-step activation
mode that promoted the formation of porous structure because
of the release of CO2, H2O, and NO2 step by step under the mild
activation conditions. Therefore, the developed CaOAAl2O3 sorbent
showed a larger pore volume (0.85 cm3/g) and higher specic surface area (958 m2/g) compared to the untreated CaO (pore volume:
0.03 cm3/g, specic surface area: 5 m2/g). Moreover, XRD patterns
exhibited the formation of the stabilizer Ca3Al2O6 in the developed
CaOAAl2O3 sorbent after 16 carbonation and calcination cycles,
which prevented the severe sintering of CaO particles. The developed CaOAl2O3 and untreated CaO sorbents showed the CO2 capture capacity of 0.35 g of CO2/g of sorbent after 100 cycles and
0.2 g of CO2/g of sorbent after 50 cycles, respectively (carbonation
at 650 C under 20% CO2 and calcination at 850 C under 100% N2).
Angeli et al. [77] developed a CaOACa3Al2O6 sorbent by a modied
solgel method. TEA (complexing agent) was added into a mixture
of Ca(NO3)24H2O and Al(NO3)39H2O diluted in distilled water at
the temperature of 50 C to obtain a molar ration of 1:1
(TEA/metals). A brown viscous solution was obtained when the
most of the water was evaporated. This solution was dried overnight at 185 C and then calcined at 900 C for 1.5 h in air. This sorbent indicated a high CO2 uptake capacity of 0.45 g of CO2/g of
sorbent (corresponding to 84% carbonation conversion), which
was maintained through 45 cycles under mild conditions (calcination at 800 C under 100% N2). The high and stable CO2 uptake
capacity was attributed to the synthesis procedure. The TEA
decomposition emitted a large amount of gases, which led to the
formation of coral-like structure. In addition, the TEA-ion complexes formed during the preparation step resulted in the uniform
dispersion of Al and Ca ions. It is worth mentioning that the

426

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

proposed sorbent represented higher stability compared to previously reported Ca-based sorbents at severe conditions. It maintained approximately 40% of its initial CO2 uptake capacity at the
end of the 100th cycle under severe conditions (calcination at
950 C under 100% CO2). Radfarnia and Sayari [78] employed a
citrate-assisted solgel technique followed by two-step calcination
procedure (under inert and air atmosphere) to develop an efcient
Al-stabilized CaO sorbent (92.5 wt.% CaO/7.5 wt.% Al2O3). After 31
carbonation/calcination cycles, the CO2 capture capacity of the synthetic sorbent was 0.57 and 0.33 g of CO2/g of sorbent under mild
and severe calcination conditions (mild condition: 800 to 900 C,
100% N2; severe condition: 930 C, 100% CO2), respectively. They
reported no loss of activity during 31 cycles, when calcination
was performed at 800 C under pure N2 ow. The high stability
of the proposed sorbent was on account of the uniform dispersion
of Ca9Al6O18 binder throughout the CaO matrix, which controlled
the structural sintering.
Ca-based sorbents derived from Ca-Al layered double hydroxides (LDHs) exhibit high stability during multiple carbonation/calcination cycles due to the uniform distribution of calcium
aluminates between CaO particles [8184]. Chang et al. [82]
applied a solgel method to develop CaAAl LDH-derived
mixed-metal oxides by using Al(OiPr)3 and Ca(NO3)2 as the precursors and hexadecyl trimethyl ammonium bromide (CTAB) as the
structure-directing agent. For the developed sorbents with a
Ca2+/Al3+ ratio of 1:1, only the Ca12Al14O33 phase was identied
after calcination at 600 C. For the sorbents with a higher
Ca2+/Al3+ ratio, the CaO phase was observed along with
Ca12Al14O33. The CaAAl mixed-metal oxides (CAMO) displayed
high specic surface areas of up to 191 m2 g1 and a pore size distribution in the range of 36 nm, allowing rapid diffusion of CO2
throughout the sorbent, inducing relatively rapid CO2 absorption
kinetics and enhancing the sintering-resistant nature through multiple carbonation/calcination cycles for CO2 capture. The CAMO
showed a high CO2 capture capacity of 49 wt.%, as well as a fast
CO2 absorption kinetics in considerably short period of 5 min for
the Ca2+/Al3+ = 7 composition. Moreover, this sorbent exhibited
highly stable CO2 capture capacity at high temperatures with only
26% capacity decay after 50 multiple carbonation/calcination
cycles. As a conclusion, the CAMO framework is a good insulator
for inhibiting the aggregation of CaO particles and therefore, it is
suitable for long-term cyclic operation at high temperatures.
Chang et al. [83] developed CaAAl LDH nanoparticles by a reverse
microemulsion method. The results showed no apparent reduction
in CO2 sorption capacity over multi-cycle carbonation/calcination
experiments because of the formation of Ca12Al14O33 oxide during
nanoparticle synthesis, which could avoid the intimate contact
between CaO nanoparticles. Yu et al. [84] employed an hydrothermal method to synthesize calciumAaluminum carbonate
(CaAAlACO3) sorbents with the Ca/Al molar ratio of 1:1 to 30:1
from calcium acetate and aluminum nitrate. By increasing the
Ca/Al molar ratio, the CO2 capture capacity increased from 13.4
to 74.2 wt.% because more Ca2+ loading in CaAAlACO3 formation
resulted in higher concentration of CaO in the synthetic sorbents
after calcination. However, the stability of the sorbent in cyclic
operation might decline with high Ca/Al ratio. The stability of the
CaAAlACO3 sorbent with the Ca/Al molar ratio of 7:1 during 10
cycles was 99% in TG analyzer (regeneration at 750 C under
100% N2) and 76% in the reactor (regeneration at 750 C under
40% CO2).
2.1.2.2. Mg-stabilized CaO. MgO is another inert support material
that is capable of stabilizing the CO2 uptake because of its high
Tammann temperature (1276 C). To obtain a very high and stable
CO2 capture capacity, it is imperative to mix the Ca and Mg ions on
a molecular level. The Mg-stabilized CaO-based sorbents that are

composed of microscopic CaCO3 and MgCO3 crystals show the loss


of CO2 capture capacity similar to natural limestone [85]. The morphology of the sorbent is inuenced by different parameters,
including the synthesis method, Ca and Mg precursor, and Ca/Mg
ratio. Li et al. [86] synthesized several Mg-stabilized CaO sorbents
using various techniques: co-precipitation, dry physical mixing,
wet physical mixing and solution mixing. Multi-cyclic carbonation/calcination experiments revealed that the most durable sorbents with high CO2 sorption capacity were prepared by the two
physical mixing methods. In addition to the synthesis method,
the MgO precursor had some effect on the performance of synthetic sorbent. However, this effect was not as strong as that of
the mixing method. CaO-based sorbent doped with MgO nanoparticles developed by thermal decomposition of magnesium oxalate
exhibited the best performance. The sorbent, which is stabilized
by 26 wt.% MgO and prepared by dry physical mixing method,
had a CO2 sorption capacity as high as 0.53 g of CO2/g of sorbent
after 50 isothermal carbonation/calcination cycles at 758 C. For
the pure CaO sorbent obtained from the same source, the initial
CO2 capture capacity of 66 wt.% decreased to 22.1 wt.% after 50
cycles under the same operating conditions. Liu et al. [87] prepared
Mg-stabilized CaO sorbents via a simple wet-mixing method from
different calcium and magnesium precursors. The best cyclic CO2
capture performance belonged to the sorbents produced from calcium and magnesium salts of D-gluconic acid, because of the uniform distribution of MgO nanoparticles among the CaO particles
in these sorbents. For instance, the sorbent containing 25 wt.%
MgO and synthesized from calcium and magnesium D-gluconate
hydrate precursors showed the CO2 capture capacity of 0.56 g of
CO2/g of sorbent over 24 cycles, which was very close to its theoretical capacity (0.59 g of CO2/g of sorbent).
Recently, Lan and Wu [88] developed different samples of
Nano-CaO/MgO-based sorbents by employing a synthesis method
consisting of three steps: (1) preparation of magnesium sol by addition of citric acid solution to MgO slurry, (2) addition of magnesium
sol to nano-CaCO3 slurry, and (3) calcination of mixture.
Characterization analyses revealed the improvement of specic
surface area (from 9.9 to 15.3 m2 g1) and average pore radius
(from 16 to 30 nm) of sorbents with the increase of MgO content,
due to the emission of a large amount of gases (CO2 and H2O) from
the decomposition of magnesium sol during precalcination process.
In addition, the increase of the precalcination temperature from
500 to 900 C led to the complete decomposition of CaCO3 and pore
structure changes while the increase of the precalcination time
caused no signicant change in the structure. According to carbonation/decarbonation experiments, the best performance belonged
to the sorbent with the nano-CaO/nano-MgO weight ratio of 3/1
that showed the higher reaction rate (by 30%) and adsorption
capacity (2-fold) in comparison with the nano-CaO/Al2O3-based
sorbent.
2.1.2.3. Zr-stabilized CaO. ZrO2 is another compound that is able to
effectively stabilize the structure of CaO sorbent. Lu et al. [45] synthesized several metal (Si, Ti, Cr, Co, Zr and Ce)-stabilized CaO sorbents by employing the ame spray pyrolysis (FSP) method.
Among all developed sorbents, Zr-stabilized CaO sorbent represented the best CO2 sorption activity under identical operating
conditions. This study showed the superiority of FSP method in
developing a sorbent with nanosize particles, high surface area
and large pore volume, as well as the role of ZrO2 in improving
the thermal stability of sorbent. In a further study, the same
research group prepared a series of CaO-based sorbents doped by
a wide range of ZrO2 loadings via the ame spray pyrolysis (FSP)
method [89]. The sorbent with a Zr/Ca molar ratio of 0.5 displayed
remarkable stability up to 1200 carbonation/calcination cycles. The
great thermal stability of developed sorbent was due to the

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

formation of well-dispersed CaZrO3 nanoparticles that prevented


the growth of CaO grains. Nevertheless, the complexity of the FSP
technique is a matter of concern for the easy production of CO2 sorbents. Radfarnia and Iliuta [90] applied the surfactant template/sonication technique to develop Zr-stabilized CaO sorbents. They
found an optimum Zr/Ca ratio of 0.303 to maximize the stability
and CO2 capture activity of the proposed sorbents. It was shown
that the formation of calcium zirconate phase during the preparation step helped preventing the structural sintering and therefore
improved the sorbent durability. The results also showed a better
CO2 capture ability of Zr-stabilized CaO sorbent in comparison
with pure CaO in severe cyclic operating conditions.
Recently, several studies were performed to investigate the
inuence of different parameters such as synthesis method,
Ca/Zr ratio, and calcination conditions on the CO2 uptake
performance of Zr-stabilized CaO sorbents. Reddy et al. [91]
prepared Zr-stabilized CaO sorbents by depositionprecipitation,
co-precipitation, ame spray pyrolysis, and solgel methods to
assess the inuence of the synthesis method on the structure and
CO2 capture activity. According to the experimental results, the
solgel-synthesized sorbent showed the best CO2 adsorption cyclic
performance. The solgel-synthesized sample had better CO2 capture capacity than ame-spray-pyrolysis-synthesized sample
because of higher Ca/Zr atomic ratio (3.57 vs. 1.5). Although
co-precipitation-synthesized and depositionprecipitation-synthe
sized sorbents possessed very high Ca/Zr atomic ratio (8.9 and
9.46, respectively), they represented signicantly lower molar
conversion compared to the solgel-synthesized and amespray-pyrolysis-synthesized sorbents due to higher crystallite size.
TEM measurements revealed that the solgel-synthesized sample
had much smaller particles (1020 nm) than co-precipitationsynthesized and depositionprecipitation-synthesized sorbents
(more than 100 nm). The authors reported that the sorbents
prepared by solgel and ame spray pyrolysis had very good
stability until 1200 carbonation/calcination cycles.
Zhao et al. [92] employed a wet chemical method to synthesize
CaZrO3-stabilized CaO sorbents with three composition ratios (10,
18, and 30 wt.% CaZrO3). Sorbent with the composition ratio of
10 wt.% CaZrO3/90 wt.% CaO represented the best CO2 carbonation/decarbonation cyclic performance in the mild conditions (carbonation at 650 C under 15% CO2 and calcination at 800 C under
air). Its CO2 capture capacity increased from 0.31 g of CO2/g of sorbent in cycle 1 to 0.37 g of CO2/g of sorbent in cycle 10 and then,
stabilized at this value over 20 cycles. This behavior
(self-reactivation) was due to a densely packed microstructure of
the as-prepared powder, which developed to a more porous structure under the rst 10 carbonation/decarbonation cycles. The best
performance in severe conditions (carbonation at 650 C under
100% CO2 and calcination at 950 C under 100% CO2) belonged to
the sorbent with the composition ratio of 30 wt.% CaZrO3/70 wt.%
CaO, which showed CO2 capture capacity decreasing from 0.36 g
of CO2/g of sorbent to 0.31 g of CO2/g of sorbent under 30 cycles.
Nanoparticles of CaZrO3 (2080 nm) dispersed within the
6200 nm CaO porous matrix were considered as the main reason
for enhancing multi-cycle stability. Broda et al. [93] prepared several ZrO2-stabilized CaO sorbents using a solgel method from different calcium precursors and zirconium (IV) propoxide and
studied the inuence of various synthesis parameters such as calcium precursor and Ca2+/Zr4+ ratio on CO2 capture. The results
showed that ZrO2-stabilized CaO sorbents derived from calcium
hydroxide or calcium acetate precursor exhibited a high surface
area and pore volume and a great CO2 capture properties.
However, the sorbents developed from calcium nitrate precursor
presented a low surface area and pore volume and a very small
CO2 capture capacity for the reason that calcium nitrate melted
during calcination and therefore, created a very coarsely grained

427

solid. Moreover, the authors mentioned that decreasing the


Ca2+/Zr4+ ratio led to the improvement of the thermal stability of
the synthetic ZrO2-stabilized CaO sorbents, as well as the reduction
of CO2 capture capacity owing to the lower amount of active CaO in
the sorbent. Among all developed sorbents, the sorbent derived
from calcium hydroxide with the Ca2+/Zr4+ ratio of 95:5 showed
the highest CO2 capture capacity of 0.34 g of CO2/g of sorbent at
the end of the 90th cycle (carbonation at 800 C under 50% CO2,
calcination at 800 C under 100% N2). This sorbent outperformed
the reference limestone by 160%. The great CO2 capture performance of the developed sorbent was on account of the uniform dispersion of calcium zirconate framework with a high Tammann
temperature (1036 C), which resulted in the minimization of thermal sintering.
2.1.3. Additional treatments
There are several additional treatments of natural (limestone)
or synthetic CaO that can be applied to improve their CO2 capture
activity in multi-cyclic carbonation/calcination processes (calcium
looping process), including hydration, preheat treatment (thermal
pretreatment), recarbonation, and chemical pretreatment
(Table 3). In the following section, these methods are discussed
in details.
2.1.3.1. Hydration. The reactivation of limestone using a hydration
process is a promising approach to improve the performance of
CaO-based sorbents in multi-cyclic carbonation/calcination operations. The hydration treatment can be classied into different
groups, including hydration treatment during carbonation
[27,94102], hydration treatment during calcination [96,97] and
separate hydration treatment on CaO/CaCO3 [97,103109], which
are determined by the stage where steam or water is introduced.
Steam addition during the carbonation step increases the carbonation conversion. This positive effect of steam on carbonation
can typically be explained by two different theories: (i) the
enhancement
of
carbonation
conversion
in
the
fast
kinetically-controlled stage by the formation of Ca(OH)2 as a transient intermediate, which its carbonation is thermodynamically
more favorable than that of CaO [99102] and (ii) the enhancement of solid-state diffusion in the calcium carbonate product
layer, which is more pronounced at lower carbonation temperature
and for more sintered sorbents [95,98].
On the other hand, there are contrary results about the effect of
steam addition during calcination. Champagne et al. [96] assessed
the reactivation of two Canadian limestones (Cadomin and
Havelock) by steam injection (up to 40 vol.%) during calcination
for 15 carbonation/calcination cycles. According to the results,
steam injection during calcination increased the carbonation reactivity of the sorbent at all concentrations of steam. In fact, steam
injection during calcination led to larger pore diameter and lower
specic surface area via enhancing particle sintering. Although
lower specic surface area decreased the sorbent carrying capacity,
larger pores lessened the diffusional resistance resulted from
CaCO3 formation at the surface, giving higher carbonation conversion. Moreover, steam injection during calcination reduced the
partial pressure of CO2 in the calciner, allowing for lower calcination temperatures. Nevertheless, Rong et al. [97] found that the
rate of sorbent activity decay was accelerated by hydration treatment during calcination, maybe because of more porosity loss in
the presence of steam during calcination. They observed that the
presence of steam during calcination (20 and 40 vol.%) resulted
in 23% decrease in CaO conversion over 10 carbonation/calcination cycles in comparison to the case without steam reactivation.
Champagne et al. [96] found that steam injection during carbonation leads to a much larger increase in carbonation reactivity
of the sorbent compared to steam injection during calcination.

428

Table 3
Summary of investigations on treated CaO sorbents.
Sorbent

Treated by

Reaction conditions
Ads.

Limestone

CaCO3
Limestone
Limestone
Limestone
Limestone
Limestone
Limestone
Limestone
Limestone
Limestone
Al2O3-stabilized CaO-based
Pelletized Al2O3-stabilized CaO-based
Ca(OH)2
Limestone
Limestone
Limestone
Limestone
Limestone
Limestone
Limestone

Acetic acid
Propionic acid
Pyroligneous acid
Citric acid
Ethanol
Steam (hydration)
Steam (hydration)
Pre-heat
Ethanol/water solution with the
volume ratio of 3
Pre-heat (950 C, 100% air, 12 h) and
recarbonation (850 C, 90% CO2,
3 min)
Steam (hydration)
Acetic acid
Steam
Steam
Acetic acid
Vinegar
Formic acid
Oxalic acid
10 wt% Aluminate cement, 10% Starch
Steam (hydration)
Steam
Steam
Pre-heat
Recarbonation after each carbonation
(800 C, 100% CO2, 5 min)
Heat-pretreated
Heat-pretreated
Heat-pretreated & recarbonation at
800 C in 90% CO2 for 3 min
Ground + Heat-pretreated (850 C,
12 h)
Ground + Heat-pretreated (950 C,
12 h)
Liquid water and steam hydration

650 C,
700 C,
700 C,
700 C,
700 C,
700 C,
780 C,
800 C,
600 C,

Number
of cycles

15% CO2, 20 min


15% CO2, 20 min
15% CO2, 30 min
15% CO2, 30 min
15% CO2, 20 min
15% CO2, 30 min
100% CO2
50% CO2, 30 min
50% CO2, 45 min

920 C,
850 C,
850 C,
750 C,
920 C,
850 C,
960 C,
800 C,
700 C,

80% CO2, 15 min


100% N2, 10 min
100% N2, 15 min
100% Ar, 30 min
80% CO2, 15 min
100% N2, 10 min
100% CO2
100% N2, 10 min
100% N2, 20 min

CO2 uptake at last


cycle (g-CO2/g-ads)

DFR
DFR
DFR
TGA
DFR
TGA
TGA
TGA
TGA

20
100
103
18
15
10
10
30
11

0.39
0.24
0.26
0.485
0.31
0.55
0.47
0.39
0.62

[121]
[120]
[122]
[123]
[118]
[104]
[105]
[110]
[119]

Ref.

650 C, 15% CO2, 5 min

850 C, 100% air, 5 min

TGA

50

0.27

[116]

650 C,
650 C,
650 C,
650 C,
650 C,
650 C,
650 C,
650 C,
700 C,
650 C,
600 C,
600 C,
650 C,
650 C,

900 C,
920 C,
925 C,
925 C,
850 C,
850 C,
850 C,
850 C,
900 C,
940 C,
900 C,
900 C,
850 C,
875 C,

TGA
TGA
TGA
TGA
TGA
TGA
TGA
TGA

FB (Fluidized-bed)
TGA
TGA
TGA
TGA

10
20
15
15
20
20
20
20
10
5
10
10
50
75

0.40
0.09
0.11
0.16
0.23
0.15
0.22
0.25
0.65
0.37
0.52
0.31
0.17
0.13

[97]
[126]
[96]
[96]
[124]
[124]
[124]
[124]
[127]
[106]
[95]
[95]
[113]
[115]

15% CO2, 25 min


15% CO2, 20 min
15% CO2, 20 min
15% CO2, 15% H2O, 20 min
15% CO2, 20 min
15% CO2, 20 min
15% CO2, 20 min
15% CO2, 20 min
15% CO2 in air, 30 min
15% CO2 in air, 15 min
20% CO2/20% H2O/60% N2, 10 min
20% CO2/20% H2O/60% N2, 10 min
15% CO2, 5 min
50 kPa CO2 in air, 5 min

100% N2, 10 min


100% CO2, 10 min
60% CO2, 15% H2O, 5 min
60% CO2, 40% H2O, 5 min
100% N2, 5 min
100% N2, 5 min
100% N2, 5 min
100% N2, 5 min
100% air, 20 min
70% CO2 in air, 20 min
20% CO2 in N2
20% CO2 in N2
100% dry air, 5 min
air, 5 min

650 C, 15% CO2, 10 min


650 C, 15% CO2, 5 min
650 C, 15% CO2, 5 min

900 C, 70% CO2, 5 min


950 C, 70% CO2, 5 min
950 C, 70% CO2, 5 min

TGA
TGA
TGA

20
20
20

0.09
0.05
0.03

[117]
[117]
[117]

650 C, 50% CO2, 30 min

850 C, dry air, 5 min

TGA

10

0.37

[114]

650 C, 50% CO2, 30 min

850 C, dry air, 5 min

TGA

10

0.45

[114]

780 C, 100% CO2, 40 min

960 C, 100% N2, 35 min

TGA

100

0.23

[109]

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

Limestone
Limestone
Limestone
Limestone
Limestone
Limestone
CaCO3
Limestone
Ca(Ac)2

Reactor
Reg.

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

They reported that although the presence of steam with the concentration of 15% for carbonation led to a 15% point increase in
conversion after 15 cycles, the presence of steam with the concentration of 40% for calcination only led to a 5% point increase after
15 cycles. However, Rong et al. [97] mentioned that steam addition
during carbonation (20 and 40 vol.%) resulted in only 23%
increase in the carbonation conversion of CaO sorbent after 10 carbonation/calcination cycles by decreasing the diffusion resistance
through the CaCO3 product layer. The poor enhancement of carbonation conversion by hydration treatment during carbonation
was justied by the limited increase of carbonation conversion in
the diffusion-control stage. In overall, steam hydration during carbonation and calcination processes have less impacts on the carbonation conversion in comparison to the separate hydration
treatment.
Introducing water or steam to CaO after regeneration is the
most important hydration approach. Calcium hydroxide
(Ca(OH)2) is produced during this hydration treatment and leads
to the agglomerate breakage because of its expanded molar volume. Therefore, separate hydration treatment after calcination
increases the specic surface area and pore volume of spent sorbent, resulting in the sorbent reactivation. This approach is briey
called steam reactivation. In an early work, Manovic and
Anthony [104] investigated the steam reactivation of spent limestone in a pressurized reactor at 200 C. According to the cyclic carbonation/calcination experiments, the reactivated limestone
showed the average CO2 capture capacity of about 0.55 g of
CO2/g of sorbent over 10 cycles, which was considerably higher
than that of the original sorbent (0.27 to 0.31 g of CO2/g of
sorbent).
Rong et al. [97] showed that different parameters such as steam
concentration, hydration temperature, and hydration frequency
affected the spent sorbent reactivation. Increasing the steam concentration during hydration treatment led to a better reactivation
performance by retaining more small pores and therefore, increasing the ultimate conversion of the fast kinetically control phase of
carbonation. However, hydration treatment at high temperatures
resulted in a lower carbonation conversion during carbonation/calcination cycles, which was on account of more severe sintering of
Ca(OH)2 at a higher hydration temperature. With respect to the
hydration frequency, hydration after every 3 cycles and once hydration did not represent satisfying reactivation performance during
cyclic CO2 sorption operation because of more severe sintering of
CaO sorbents derived from Ca(OH)2. The activity of the spent sorbent was recovered by separate steam hydration after every calcination step. Recently, Coppola et al. [106] assessed the effects of
water hydration on the CO2 capture capacity and attrition tendency
of a limestone-derived CaO sorbent. The results demonstrated that
the CO2 capture capacity of sorbent increased from 0.04 g of CO2/g
of sorbent in the last carbonation before hydration to 0.320.37 g of
CO2/g of sorbent in the rst carbonation after hydration, because of
particle swelling and development of active porosity. However, the
capacity quickly decayed along with the cycles due to the severe
sintering of Ca(OH)2-derived CaO sorbents. In addition, the sorbent
hydrated for 60 min showed larger CO2 capture capacity, as well as
limited attrition tendency. In fact, two distinct time scales should
be considered for the optimal design of hydration stage: (i) the time
required for full hydration of the free lime (water uptake) and (ii)
the time required for improving the connectivity and mechanical
stability of particles by wet chemical sintering. Although 10 min
was enough for water uptake, wet chemical sintering was not
achieved during this time. Moreover, more soaking time (more than
60 min) led to a decrease of CO2 uptake capacity without any significant improvement in its attrition resistance.
Hydration treatment of the spent limestone not only recovers
its CO2 capture capacity but also signicantly increases its attrition

429

tendency, which restricts the industrial applicability of this


method. The cost impact related to the production of steam is also
another matter of concern for a high level of hydration. The partial
hydration of limestone is proposed as an acceptable strategy in
order to decrease both negative impact on the mechanical strength
of reactivated material and steam consumption. In addition, pelletization can be another alternative to improve the mechanical
strength of sorbents [95,107,108].
2.1.3.2. Thermal pretreatment and recarbonation. Thermal pretreatment of limestone has been proposed as another activation
approach, which results in the stabilization of material structure
and self-activation of sorbent during cyclic carbonation/calcination
operations (increase of sorbent activity along with the cycle number). Manovic and Anthony [110] examined the thermal pretreatment of four Canadian limestones. According to the CO2 sorption
experiments, pretreated sorbents showed better conversions at
the end of cyclic operation in comparison to the corresponding
original sorbents (natural limestones). After 30 cycles, a CO2 sorption capacity up to 0.39 g of CO2/g of sorbent was obtained for the
studied limestones. The authors also proposed a poreskeleton
model to explain self-activation phenomenon (Fig. 4). They
demonstrated that the formation of hard skeletons during thermal
pretreatment of a sorbent stabilizes its structure. The transformation of the hard skeleton to a soft skeleton during cyclic operation
increases the CO2 sorption activity of sorbent by facilitating the
mass transfer. In fact, the hard skeleton formed during the thermal
preatreatment of CaO sorbent is less active and more stable in
comparison to soft skeleton and therefore, results in a lower CO2
sorption capacity in the initial cycles. The continuous growth of
soft skeleton and decline of hard structure leads to the enhancement of sorbent activity until the stabilization takes place between
skeleton changeovers. In a further study, Chen et al. [111] also
observed the self-activation of Strassburg limestone and Arctic
dolomite thermally pretreated at 1000 and 1100 C, respectively,
over 1000 carbonation/calcination cycles. However, as mentioned
by arias et al. [112], the self-activation may not be effective under
typical conditions of a circulating uidized-bed carbonator, where
the reaction time is restricted to few minutes at low CO2 partial
pressures (below 10 kPa) and temperatures around 650 C.
Recently, it was found that conversion of thermal pretreated
CaO sorbent was based on the balance between the increase of surface area because of enhanced solid-state diffusion carbonation
and the decrease of surface area of renovated CaO structure due
to sintering during calcination stage [113,114]. Heat pretreatment
led to an extremely sintered CaO structure, which decreased carbonation conversion in the fast kinetically controlled phase and
increased carbonation conversion in the solid-state diffusion

Fig. 4. Schematic representation of poreskeleton model [110].

430

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

controlled phase in the 1th cycle. Upon calcination, carbonation


conversion in the fast kinetically controlled phase of the 2nd cycle
was increased because of the enhancement of surface area of the
regenerated CaO structure. However, the carbonation conversion
in both kinetically and solid-state diffusion controlled phases
was decreased at a small rate after the 2nd cycle because of the sintering of the regenerated CaO structure during calcination process
at high temperatures [114]. Valverde et al. [113] showed that
harshening calcination conditions (calcination temperatures and
time periods) precluded self-reactivation of thermal pretreated
CaO sorbent by speeding up the renovated soft skeleton sintering,
which resulted in a conversion reduction in the fast kinetically
controlled phase. According to their results, self-reactivation of
the proposed sorbent completely hindered at the calcination temperatures above 900 C (time period: 5 min) or for calcination time
period of 15 min (temperature: 850 C). Sanchez-Jimenez et al.
[114] found that sorbent grinding prior to heat pretreatment
resulted in a slow decay rate of carbonation in the diffusion controlled phase and so, a steady increase of conversion in the fast
kinetically controlled phase with the cycle number. The experimental results obtained for raw and preground limestone (without
thermal pretreatment) revealed that the carbonation conversion in
the solid-state diffusion controlled phase was signicantly
enhanced by pregrinding. This enhancement was attributed to
local high stresses induced by grinding, which exceeded the cohesion forces between the lattice atoms, and therefore resulted in
crystal cracking and enhancement of structural defects that helped
CO2 diffuse better through the solid. The authors also reported that
heat pretreatment in a CO2 atmosphere led to the inhibition of carbonation in the solid-state diffusion controlled phase and a high
increase of the carbonation rate in the fast kinetically-controlled
phase, which could be ascribed to annealing of the crystal
structure.
The introduction of a recarbonation step between carbonation
and calcination steps is a novel process aimed to increase the
CO2 capture capacity of CaO-based sorbents in multicyclic CO2 capture systems. Grasa et al. [115] studied the kinetics of the carbonation reaction in recarbonation step under the conditions of high
temperature and CO2 partial pressure. The experimental results
showed that the addition of a short recarbonation step (100
200 s) on partially carbonated CaO sorbents stabilized the sorbent
capture capacities at 0.150.20 M conversion. The carbonation
reaction in recarbonation step, which mainly occurred in the slow
diffusion controlled reaction phase, was strongly favored by high
recarbonation temperatures (750800 C), high CO2 partial pressures (beyond 60 kPa), and a certain presence of steam.
Valverde et al. [116] indicated that the synergetic combination
of heat pretreatment and recarbonation resulted in a high and
stable CaO conversion in the carbonation step of multiple carbona
tion/recarbonation/calcination cycles. To this end, a natural limestone was preheated at 950 C for 12 h in a dry air atmosphere
and then, was subjected to the multi-cyclic carbonation/calcination and carbonation/recarbonation/calcination experiments. The
experimental results revealed that the heat pretreated sorbent
showed a stable but very small CaO conversion through carbonation/calcination multi-cyclic operations. Indeed, carbonation under
low CO2 concentrations and calcination at temperatures above
850 C prevented the reactivation of sorbent by heat pretreatment.
However, the proposed sorbent indicated a stable and high CaO
conversion from the 2nd cycle in the multi-cyclic carbonation/re
carbonation/calcination experiments. In fact, the addition of a
recarbonation step between the carbonation and calcination steps
led to a reactive and thermally stable CaO structure after calcination by the signicant enhancement of solid-state diffusion during
recarbonation. Later, the same research group studied the inuence of the same reactivation methods (heat pretreatment and

recarbonation) on the multi-cyclic CO2 capture performance of


limestone derived CaO sorbent for the cases with a high CO2 partial
pressure in the calciner [117]. The authors claimed that the results
were in contrast with those reported in their previous work [116]
and the inuence of recarbonation on the cyclic CO2 sorption was
affected by the CO2 partial pressure in the calciner. When the CO2
partial pressure during calcination was far enough from the equilibrium pressure, the irreversible desorption of CO2 governed
decarbonation, which resulted in a highly porous CaO structure
with increased surface area for the fast kinetically-controlled carbonation phase. Addition of a recarbonation stage prior to calcination led to the more enhancement of porosity in the resultant CaO
structure, because it let the following decarbonation happen deeper in the bulk of solid. It was also mentioned that the heat pretreatment promoted the desirable effect of recarbonation.
However, during calcination at high CO2 partial pressure, decarbonation was ruled by a dynamic and reversible CO2 adsorption/desorption mechanism, which was precluded by the
addition of a recarbonation stage before calcination. In addition,
this mechanism precluded the growth of the CaO crystal structure
and therefore, prevented the carbonation in the fast
kinetically-controlled phase. On the contrary, the heat pretreatment alone showed favorable effects on the cyclic CO2 capture process under harsh calcination conditions. Heat pretreatment and
calcination under high CO2 partial pressure led to the enhancement
of CaO conversion during the diffusion-controlled phase. Heat pretreatment allowed also decreasing the calcination temperature at
high CO2 partial pressure.
2.1.3.3. Chemical pretreatment. Although the cost impacts associated with the use of chemical solutions can be a matter of concern,
treatment by chemical solutions is another promising approach to
enhance the CO2 capture activity of CaO sorbent. Li et al. [118]
evaluated the CO2 uptake performance of a limestone derived
CaO sorbent treated with 50%, 70% and 90% ethanol/water solutions. They found that the CO2 capture capacity of the sorbent treated with ethanol/water solution was higher than that of sorbent
hydrated with distilled water and much higher than that of sorbent
derived directly from limestone. Higher concentration of ethanol in
solution led to higher CO2 sorption capacity and better
anti-sintering performance for treated CaO sorbent. The enhanced
CO2 adsorption capacity of modied sorbent was attributed to the
fact that the ethanol molecule increases H2O molecule afnity and
penetrability to CaO, resulting in the higher specic surface area
and larger pore volume after calcination. In a similar work, Wang
et al. [119] treated CaO sorbents derived from calcium acetate with
ethanol/water solution at different temperatures. They concluded
that CaO modied by ethanol/water solution (volume ratio of 3)
at room temperature had a higher CO2 capacity and better stability.
For this adsorbent the authors reported CO2 sorption capacity of
74 wt.% in the rst cycle and 62.5 wt.% in the eleventh cycle. In fact,
the addition of ethanol to water decreased the solute solubility of
mixture and resulted in smaller particle size, larger surface area
and pore volume and, consequently, higher capacity. However,
the treatment with pure ethanol or pure water led to poor capacity.
The reaction of limestone with organic acids is known as
another chemical treatment that improves the sintering resistance
of CaO sorbents derived from limestone by altering the porous
structure [120124]. Li et al. [121] investigated the modication
of limestone with 50% acetic acid solution (the molar ratio of acetic
acid to calcium: 1.5:1). The modied limestone exhibited the sorption capacity of 0.39 g of CO2/g of sorbent after 20 cycles, which
was signicantly higher than that corresponding to the natural
limestone (0.12 g of CO2/g of sorbent). The better stability of acidied limestone during cyclic operation was attributed to its smaller
particle sizes, higher surface area and pore volume obtained during

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

acidication. These results were conrmed by Li et al. [122] who


reported an enhancement in cyclic stability of limestone treated
with pyroligneous acid (PA) (ratio of PA to limestone: 20 mL/g).
They also mentioned that the cost of PA is lower and hence, the
use of PA is more economical compared to acetic acid. With the
purpose of reducing the cost of sorbent production, natural limestone was also considered by Radfarnia and Iliuta [123] and a novel
synthesis technique (limestone acidication by citric acid followed
by two-step calcination in Ar and air atmospheres) was applied in
order to prepare highly porous CaO structure with unique CO2 capture ability. The performance of the proposed sorbent was investigated in detail, revealing a much better stability and CO2 sorption
activity of the developed sorbent compared to natural limestone.
The principal of the developed synthesis technique consisted in
the decomposition of calcium citrate (product of limestone acidication) in controlled atmosphere (Ar) to produce in-situ carbon,
which controlled the particle size enlargement. The carbon combustion during the secondary calcination step (in air) promoted
the dispersion of particle agglomerates and enhanced the material
porosity. Ridha et al. [124] proposed the treatment of limestone by
four organic acids with the aim of improving the cyclic CO2 capture
performance of limestone. The increase of CO2 capture capacity for
the modied sorbents was explained by the fact that the acidication by organic acids widened the pores of the modied sorbents
and improved the resistivity of them to sintering phenomenon.
However, the authors claimed that the treatment of limestone with
organic acids has little positive effect on CO2 capture capacity of
sorbents, but contributes in great measure to the process cost. In
addition, the reactivity of modied sorbents towards SO2 was
improved, resulting in the acceleration of the reduction of their
CO2 capture capacity compared to untreated limestone.
Many treatments used to improve the cyclic CO2 capture behavior of CaO-based sorbents, such as acidication and hydration, have
the drawback of decreasing the mechanical strength of sorbents. In
such cases, pelletization is often considered as an option to
improve the mechanical strength of sorbents [61,125]. However,
it should be noted that pelletization is an expensive process and
therefore, the application of a pretreated sorbent is considered economical when it shows a high improvement of CO2 sorption activity [126]. Ridha et al. [126] determined the inuence of
acidication and pelletization on CO2 capture behavior of
CaO-based pellets. For this purpose, raw limestone was treated
by acetic acid solution (10 vol.%) and then, both untreated and
acidied limestones were pelletized by a calcium aluminate
cement binder (1014 wt.%). According to the experimental
results, the acidied pellets adsorbed 41% more CO2 than
un-acidied pellets after 20 repetitive cycles of carbonation
(650 C, 15% CO2) and calcination (920 C, 100% CO2), because of
the enhanced morphology of acidied pellets. Indeed, the acidied
pellets possessed large pores of size within 20200 nm, which
enabled the sorbent to continue hosting more CaCO3. However,
the increase in CO2 capture capacity could not justify such treatments due to the high price of acetic acid (around $900/ton).
Moreover, the acidied pellets also showed an improved reactivity
towards SO2. A pretreatment of ue gases to remove SO2 is therefore necessary when acidied pellets are used for CO2 capture.
Chemical pretreatment is not limited to treatment with ethanol/water solution and acidication. Chen et al. [127] studied the
effect of pelletization of sorbent and addition of pore forming
agents on the attrition resistance and CO2 capture capacity of
Ca-based sorbent. The original limestone, aluminate cement (consisting of 58 wt.% Al2O3) and starch were used as precursors. The
experimental results demonstrated that pelletization with
10 wt.% aluminate cement considerably improved the mechanical
property of pellets, because of the high mechanical strength of aluminate. Additionally, the pellets developed with 10 wt.% aluminate

431

cement and 510 wt.% starch indicated a higher and more stable
CO2 capture capacity in multi-cyclic carbonation/calcination
experiments in comparison to the natural limestone. After 10 carbonation (at 700 C in 15% CO2/air balance) and calcination (at
900 C in air) cycles, the pellets containing 10 wt.% aluminate
and 10 wt.% starch offered a capture capacity of 0.65 g of CO2/g
of pellets, which was higher compared to pellets of natural limestone (0.5 g of CO2/g of pellets). The excellent performance of
the pellets developed with aluminate cement and starch was
attributed to the increase of pore volume of pellets because of
starch decomposition during calcination and the deceleration of
sintering due to the presence of alumina which has a high melting
point.
2.2. Ceramic CO2 sorbents
In addition to CaO-based sorbents, alkaline ceramic materials
have also been proposed as potential candidates for CO2 removal
processes such as SESR [4,6,128137]. However, their kinetic limitations during the CO2 capture still remain their main hurdle. The
literature data for lithium zirconate (Li2ZrO3), lithium orthosilicate
(Li4SiO4), and sodium zirconate (Na2ZrO3), which are the most
investigated ceramic sorbents, is summarized in Table 4. The carbonation reactions for the most well-known alkaline oxide sorbents are as follow:

Li2 ZrO3 s CO2 g $ Li2 CO3 s ZrO2 s DH 298 160 kJ=mol
2
Na2 ZrO3 s CO2 g $ Na2 CO3 s ZrO2 s DH 298 149 kJ=mol
3
Li4 SiO4 s CO2 g $ Li2 CO3 s Li2 SiO3 s DH 298 143 kJ=mol
4
2.2.1. Lithium zirconate (Li2ZrO3)
In 1998, Nakagawa and Ohashi [4] proposed Li2ZrO3 as a
promising candidate for CO2 adsorption. The material has a theoretical uptake capacity of 0.28 g of CO2/g of sorbent in the temperature range of 450600 C. They employed solid-state reaction of
Li2CO3 and ZrO2 precursors to prepare Li2ZrO3. The synthesized
sorbent showed the CO2 capture capacity of 0.22 g of CO2/g of sorbent at the temperature of 500 C under a ow containing 20 vol.%
CO2 and 80 vol.% H2. In a further study, the same research group
[135] synthesized potassium carbonate doped Li2ZrO3 (K-Li2ZrO3)
via solid-state reaction method. The developed sorbent showed a
higher CO2 adsorption rate in comparison to the undoped
Li2ZrO3. The signicant improvement in kinetic rate was ascribed
to the formation of a eutectic molten carbonate layer of Li2CO3
and K2CO3 above 500 C, which considerably increased the CO2 diffusion rate. Later, Ida and Lin [136] proposed a comprehensive
double-shell model to explain the mechanism of CO2 chemisorption on pure and potassium-doped Li2ZrO3 sorbents. According to
their model, CO2 reacts with Li+ and O2 ions after diffusion to
the surface of Li2ZrO3 to form ZrO2 and Li2CO3. These materials
form a double solid shell around the unreacted Li2ZrO3 (Fig. 5).
At this step, Li+ and O2 have to diffuse through the ZrO2 layer to
react with CO2 and also CO2 molecules have to diffuse through
the Li2CO3 layer for reaction. Therefore, the adsorption rate begins
to decrease. In the case of K-Li2ZrO3, the formation of a eutectic
molten carbonate layer of Li2CO3 and K2CO3 enhances the adsorption kinetic rate by increasing the CO2 diffusion rate through the
external layer. However, Ochoa-Fernandez et al. [138] showed that
despite the positive effect of potassium carbonate doping into
Li2ZrO3 on the kinetic rate of CO2 adsorption, the CO2 capture

432

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

Table 4
Summary of investigations on ceramic-type sorbents.
Sorbent

Synthesis technique

Reaction conditions

Li2ZrO3
Li2ZrO3
Li2ZrO3
Li2.2ZrO3.1
KALi2ZrO3a
K0.2ALi1.6ZrO2.9
K0.2ALi1.6ZrO2.9
KALi2ZrO3b
Li2ZrO3

Citrate solgel
Co-precipitation
Liquid-state
Liquid-state
Solid-state
Citrate solgel
Liquid-state
Solid-state
Ultrasound-assisted
surfactant-template
Solid-state

Solid-state
Solid-state
Solid-state
Solid-state
Solid-state
Liquid-state
Liquid-state
Surfactant template/
ultrasound assisted

500 C,
550 C,
575 C,
575 C,
550 C,
550 C,
575 C,
500 C,
575 C,

50% CO2, 60 min


100% CO2, 20 min
100% CO2, 12 min
100% CO2, 20 min
60% CO2, 100 min
25% CO2, 60 min
100% CO2, 20 min
100% CO2, 190 min
100% CO2, 30 min

650 C,
690 C,
650 C,
630 C,
800 C,
675 C,
630 C,
750 C,
690 C,

100%
100%
100%
100%
100%
100%
100%
100%
100%

525 C,
580 C,
580 C,
680 C,
550 C,
550 C,
700 C,
600 C,
550 C,
575 C,
575 C,

15% CO2, 60 min


4% CO2, 60 min
4% CO2, 60 min
100% CO2, 15 min
100% CO2, 90 min
100% CO2, 90 min
50% CO2, 30 min
100% CO2, 30 min
100% CO2, 25 min
50% CO2
100% CO2, 10 min

850 C,
700 C,
700 C,
800 C,
550 C,
550 C,
700 C,
780 C,
800 C,
680 C,
840 C,

60 min
100% N2, 15 min
100% N2, 15 min
100% N2, 10 min
100% N2, 90 min
100% N2, 90 min
100% N2, 30 min
100% N2, 65 min
100% N2, 55 min
100% Ar
100% Ar, 30 min

Ads.

Li2CO3/K2CO3-doped Li2ZrO3
KALi4SiO4
NaALi4SiO4
Li4SiO4
Li4SiO4
Ball milled Li4SiO4
Li4SiO4
Na2ZrO3
Na2ZrO3
Na2ZrO3
Na2ZrO3
a
b
c
d
e

Reactor

Number
of cycles

TGA
TGA
TEOM
TEOM
TGA
TGA
TEOMc
TGA
IGAd

3
3
100
8
6
4
8
5
11

0.26
0.23
0.24
0.24
0.2
0.23
0.18
0.26
0.22

[143]
[134]
[5]
[138]
[130]
[144]
[138]
[142]
[145]

TGA
TGA
TGA
TFBe
TGA
TGA
TGA
TGA
TGA
TEOM
IGA

12
25
25
15
10
10
16
2
2
8
4

0.083
0.15
0.07
0.3
0.044
0.154
0.28
0.24
0.15
0.15
0.13

[139]
[150]
[150]
[149]
[147]
[147]
[148]
[6]
[207]
[7]
[156]

Reg.
N2, 50 min
N2, 10 min
Ar
N2
N2, 30 min
N2, 50 min
N2
N2, 230 min
Ar, 30 min

CO2 uptake at last


cycle (g-CO2/g-ads)

Ref.

Li2CO3:ZrO2:K2CO3 (1.15:1.0:0.2).
91.3% Li2ZrO3 + 3.4% Y2O3 + 0.2% Al2O3 + 5.1% K2O.
Tapered Element Oscillating Microbalance.
Intelligent Gravimetric Analyzer.
Twin Fixed-Bed Reactor.

Fig. 5. Adsorption mechanism on Li2ZrO3 solid sorbent [130].

capacity and cyclic stability of sorbent was lowered by the addition


of potassium.
Alkali zirconates are normally prepared by solid-state reaction
of ZrO2 and alkali salts such as alkali carbonate
[4,129,130,135,136,139]. The synthesized materials often show a
slow kinetic rate and a small capture capacity because of their
large particle sizes and poor porosity. Hence, the CO2 capture properties of alkali zirconates can be improved by the reduction of their
particle sizes, which can be achieved by controlling the starting
precursor sizes. For instance, Xiong et al. [140] studied the preparation of K-doped Li2ZrO3 by the solid-state reaction of ZrO2 (with
two different precursor sizes of 1 and 45 lm), Li2CO3, and K2CO3.
They found that the particle size of developed sorbent was
decreased by using ZrO2 of smaller particle size. The authors also
observed that the sorbent with smaller particle size had higher
CO2 adsorption rate. However, it should be noted that the nal product (Li2ZrO3) is subsequently prepared by the solid-state reaction
of mixed powders of ZrO2 and Li2CO3 at high temperature, which
results in material sintering. Therefore, reducing the particle size
of sorbent with decreasing the particle sizes of starting materials
is a problematical issue.
Some researchers tried to improve the CO2 uptake properties of
Li2ZrO3 by developing different synthesis methods [5,134,141
145]. Ochoa-Fernandez et al. [5] prepared nanocrystalline tetragonal and monoclinic phases of Li2ZrO3 by a novel liquid-state

soft-chemistry route from zirconoxy nitrate and lithium acetate


as precursors. The nanocrystalline tetragonal Li2ZrO3 developed
by the proposed method showed a faster CO2 adsorption and
regeneration rate in comparison to the sorbents synthesized by
the solid-state reaction method. This behavior was attributed to
the small crystallite size (about 13 nm) of the developed sorbent.
In addition, this sorbent maintained above 90% of its initial CO2
capture capacity after 100 adsorption/regeneration cycles. Yi
et al. [134] compared the CO2 capture performance of Li2ZrO3 synthesized via a liquid-state precipitation technique and a solid-state
reaction technique. The Li2ZrO3 sample developed by the precipitation method showed a CO2 sorption rate ten times faster than that
developed by solid-state reaction method. This observation was
attributed to the smaller particle size of the precipitated Li2ZrO3,
which led to a higher CO2 diffusion rate. The cyclic adsorption/desorption experiments showed that the CO2 capture rate and capacity were almost unchanged through three cycles. Moreover, the
authors indicated that the addition of more than 20% steam
enhanced the CO2 sorption rate considerably. In another study,
Iwan et al. [142] put forward a modied solid-state synthesis
method for the production of Li2ZrO3 CO2 sorbents. The main difference between the proposed synthesis method and the traditional solid-state reaction method was the use of zirconium
hydroxide instead of zirconium oxide at the stage of mixing with
lithium carbonate. This permitted a much closer mixing of the precursors in the aqueous slurry phase and a lower calcination temperature (700750 C) compared to that of the traditional
solid-state method (P900 C). Therefore, the agglomeration of
the nal product was prevented and a sorbent with higher surface
area (about 11 m2/g) and reactivity was produced. The CO2 capture
experiments showed that the sorbent with higher surface area,
even undoped with promoters such as potassium carbonate, possessed much higher kinetic rates. However, the adsorption rate of
the sorbent developed by the proposed synthesis method was
lower than that of tetragonal Li2ZrO3 prepared by using the
liquid-state soft-chemistry route [5].

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

Xiao et al. [143] proposed a citrate solgel method to synthesize


Li2ZrO3 nanocrystals. The nanocrystalline Li2ZrO3 with a tetragonal
phase showed a better CO2 capture performance in comparison to
the sorbents fabricated by co-precipitation or liquid-phase methods. It exhibited a faster sorption rate and a higher, nearly stoichiometric capture capacity (0.26 g of CO2/g of sorbent). Furthermore,
the developed sorbent demonstrated a good stability in cyclic
adsorption/desorption process, maintaining its initial capture
capacity (0.26 g of CO2/g of sorbent) after 3 cycles. Later, the
same research group employed a citrate solgel method to prepare
K-doped Li2ZrO3 sorbents [144]. The developed K-doped Li2ZrO3
sorbents showed better CO2 capture performance compared to
the Li2ZrO3 synthesized by a similar method, especially at low
CO2 partial pressures. The K-doped Li2ZrO3 with an optimized
K:Li:Zr molar ratio of 0.2:1.6:1 exhibited a CO2 sorption rate of
1.5 wt.%/min at 550 C and CO2 partial pressure of 0.25 bar.
Moreover, this sorbent displayed a good stability and maintained
its initial CO2 capture capacity (0.23 g of CO2/g of sorbent) after
4 adsorption/desorption cycles .In another study, Radfarnia and
Iliuta [145] developed Li2ZrO3 by applying surfactant template/sonication method and showed an increase of the CO2 capture
capacity of developed sorbent compared to the material prepared
by simple surfactant template method (without sonication) or conventional wet-mixing route. Li2ZrO3 prepared by surfactant template/sonication method contained less agglomerated structure
and its porous framework facilitated gas and ion diffusions to/from
particle layers. The sonication time and surfactant concentration
were found to affect the sorbent properties (crystallite size and
BET surface area). Porous Li2ZrO3 prepared by less surfactant and
irradiation time could achieve maximum uptake capacity of
22 wt.%, which was signicantly higher compared to other tested
samples. Despite the good sorption capacity, it was mentioned that
porous Li2ZrO3 still suffered from slow kinetics of sorption at low
CO2 partial pressure (below 0.75 bar), which can limit its application for SESR operation.
2.2.2. Lithium orthosilicate (Li4SiO4)
Lithium orthosilicate (Li4SiO4) was introduced as a novel CO2
sorbent by Kato and Nakagawa [8]. They employed the
solid-state reaction method to synthesize Li4SiO4 from Li2CO3
and SiO2 precursors. The developed sorbent displayed an adequate
CO2 sorption capacity up to 0.36 g of CO2/g of sorbent in the temperature range of 450700 C. Later, Kato et al. [131] reported
about the fascinating characters of Li4SiO4 as a CO2 sorbent.
According to their results, the CO2 sorption capacity and rate of
Li4SiO4 were, respectively, around 50% more and 30 times faster
than that of Li2ZrO3, even at CO2 concentration as low as 20%. It
was also found that at 2% CO2 concentration, Li4SiO4 clearly
adsorbed CO2 even though the rate of adsorption was less than that
at 20% CO2 concentration. Nevertheless, Li2ZrO3 showed no obvious CO2 sorption. However, Rodriguez-Mosqueda and Pfeiffer and
Yi et al. [9,134] observed that Li4SiO4 had slow kinetic rate at
low CO2 partial pressures in comparison to other alkaline ceramics
such as Li2ZrO3 and Na2ZrO3 developed by liquidliquid state
reaction.
Venegas et al. [146] studied the CO2 capture properties of
Li4SiO4 with different particle sizes. Different synthesis methods,
including solid-state reaction, precipitation and solgel, were
employed to synthesize Li4SiO4 samples with different particle
sizes. The experimental results revealed that the Li4SiO4 sample
prepared via precipitation method had the highest CO2 adsorption
reactivity. This behavior was attributed to its smaller particle size
(about 3 lm), which was associated with the presence of more
active lithium atoms over the surface of particles. Romero-Ibarra
et al. [147] initially synthesized lithium orthosilicate (Li4SiO4) by
the solid-state reaction of LiOH and SiO2 precursors, and then

433

modied it by the ball milling process. The characterization analysis of both Li4SiO4 sample and ball milled Li4SiO4 sample showed
that the crystal size was decreased from >500 to 175 and the
surface area was increased from 0.4 to 4.9 m2 /g. The modied
Li4SiO4 sample showed better efciencies during the CO2 chemi
sorptiondesorption process without further sintering effects.
Shan et al. [148] prepared Li4SiO4-based sorbents from diatomite (as an inexpensive source of silicon) by the solid-state reaction method at 700 C. They studied the effect of different molar
ratios of raw materials on CO2 adsorption capacity in a gas mixture
of CO2 and N2 (50 vol.%). They found that the CO2 adsorption capacity reached the largest value (30.32 wt.%) at Li2CO3/SiO2 molar
ratio of 2.6. They also observed that the CO2 adsorption capacity
decreased by only 6.44 wt.% during 16 adsorptiondesorption
cycles because of the specic morphologies of Li4SiO4-based sorbents synthesized from diatomite. Wang et al. [149] developed
Li4SiO4-based sorbents from Li2CO3 and different types of silica
(citric acid pretreatment rice husk ash (CRHA), nano-structured
Aerosil, and crystalline Quartz powders) by the solid-state reaction
method and studied the effects of the type of silica on the
microstructure and CO2 capture performance of developed sorbents. Among all developed sorbents, CRHA-Li4SiO4 sorbent
showed the best performance with the highest CO2 capture capacity of 30.5 wt.% more rapid adsorption/desorption process and better regenerability during multicyclic adsorption/desorption
experiments (only 2.1 wt.% decrease of CO2 capture capacity after
15 cycles). The excellent performance of this sorbent was attributed to its lower crystalline of pure Li4SiO4, smaller particle size, and
larger specic surface area, as a result of the favorable microstructure of nanoparticles and the strong sintering-resistant character of
CRHA.
Seggiani et al. [150] applied the solid-state reaction method to
synthesize Li4SiO4-based sorbents containing 10, 20, 30 wt.% of
alkali carbonates (K2CO3, Na2CO3), binary (K2CO3/Li2CO3,
Na2CO3/Li2CO3) and ternary (K2CO3/Na2CO3/Li2CO3) eutectic carbonate mixtures. The CO2 adsorption characteristics of the developed sorbents were investigated at high temperatures in the
range of 500600 C and low CO2 partial pressure of 0.04 atm.
According to the results, the CO2 sorption rate and capacity of all
the promoted Li4SiO4-based sorbents were obviously improved in
comparison to pure Li4SiO4 sorbent. This was attributed to the
much faster diffusion of CO2 through the molten carbonate shell,
compared to that through the solid Li2CO3 shell in no-promoted
Li4SiO4. Furthermore, the CO2 sorption rate and capacity of the promoted sorbents containing alkali carbonates were higher than
those of the promoted sorbents containing binary or ternary eutectic carbonate mixtures. For the promoted sorbents containing
alkali carbonates, (K/Li)CO3 and (Na/Li)CO3 eutectic melt were
formed on the Li4SiO4 surface during the CO2 sorption. However,
for the promoted sorbents containing binary and ternary eutectic
carbonate mixtures, a eutectic liquid phase was already present
on the surface of Li4SiO4 at the beginning of CO2 sorption and
CO2 had to diffuse through this molten layer in order to reach
the surface of Li4SiO4. Therefore, this molten layer acted as an additional diffusional resistance. For all the promoted Li4SiO4-based
sorbents, 580 C was reported as optimum sorption temperature,
because the equilibrium temperature was around 590 C at CO2
partial pressure of 4.04 atm, and therefore, the desorption process
was activated at temperatures higher than 580 C. At 580 C, promoted Li4SiO4-based sorbents containing 30 wt.% of K2CO3 or
Na2CO3 indicated the best CO2 adsorption characteristics with
sorption capacity of 0.23 g of CO2/g of sorbent corresponding to a
Li4SiO4 conversion of 80%. However, the promoted sorbent containing Na2CO3 indicated a considerable degeneration of CO2 capture capacity during multi-cyclic adsorption/regeneration
experiments due to the structural sintering. On the contrary, the

434

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

promoted sorbent containing K2CO3 presented a very good cyclic


stability during 25 adsorption/regeneration cycles.

2.2.3. Sodium zirconate (Na2ZrO3)


Na2ZrO3 has been proposed as an alternative for CO2 capture
because of its higher adsorption kinetic rate at high temperatures
and lower initial precursor costs in comparison to Li2ZrO3 and
Li4SiO4. Lopez et al. [6] examined the CO2 capture properties of different sorbents, including Li2ZrO3, Na2ZrO3, Li4SiO4, Na2TiO3 and
Na3SbO4, developed by solid-state reaction of precursors.
Although Na2ZrO3 had the best adsorption kinetics at high temperature compared with the others, its regeneration performance was
unfavorable compared to Li2ZrO3 and Li4SiO4.
In order to overcome the drawbacks of Na2ZrO3 (such as the
need of the high regeneration temperature and the inherent weakness of the regeneration kinetics), different research groups tried to
promote its CO2 capture performance by the incorporation of various active metals into its structure and/or the production of smaller nal particle sizes [151155]. Most of them applied the
solid-state reaction method to synthesize their sorbents.
However, the results were not fully satisfactory. Therefore, different other synthesis methods were further employed for the synthesis of Na2ZrO3 to improve its performance. Zhao et al. [7]
applied a novel soft-chemical route to synthesize nanocrystalline
Na2ZrO3 from zirconoxy nitrate and sodium citrate. During the
synthesis process, an amorphous zirconium complex was formed
and then, it was calcined in a controlled atmosphere. The strong
reaction between nitrate and citrate during calcination resulted
in the in-situ carbon formation. Subsequent carbon burnoff promoted the formation of open pore structure. In the proposed synthesis method, the crystal phase of Na2ZrO3 was controlled by the
calcination temperature and atmosphere. A two-step calcination at
800 C caused to the formation of pure monoclinic phase, which
was much more active rather than hexagonal phase of Na2ZrO3.
According to the CO2 capture studies, the monoclinic Na2ZrO3
showed much faster CO2 adsorption rates in comparison to its
hexagonal counterpart, even at CO2 partial pressure as low as
0.025 bar. It also exhibited an excellent stability over 8 adsorption/desorption cycles. Taking into consideration the improvement
of Li2ZrO3 sorption properties by the application of surfactant template/sonication technique, Radfarnia and Iliuta [156] applied the
same method to develop porous Na2ZrO3 because this material
was supposed to offer better CO2 sorption kinetics compared to
Li2ZrO3. The behavior of the proposed Na2ZrO3 sorbent was compared with that of samples prepared by surfactant template
method (without sonication) and conventional wet-mixing route.
The performance of the new developed Na2ZrO3 was unexpected.
The samples prepared by surfactant template/sonication technique
were found to be less active during cyclic operation compared to
the conventional Na2ZrO3. This behavior was interpreted by the
low resistivity of the pore structure at the high temperature treatment required during calcination (compared to Li2ZrO3), resulting
in the loss of material main porosity and the creation of agglomerated particles.
Martnez-dlCruz and Pfeiffer [157] evaluated the microstructural evolution of Na2CO3AZrO2 external shell produced during
the CO2 adsorption on Na2ZrO3 as a function of sorption temperature. According to the results, the microstructural properties varied
with the sorption temperature. At T 6 550 C, CO2 adsorption on
Na2ZrO3 was not limited because the Na2CO3AZrO2 external shell
was mesoporous. CO2 diffused through the mesopores and the
reaction continued. However, the mesoporosity of the
Na2CO3AZrO2 external shell disappeared at T > 550 C due to thermal sintering and the CO2 sorption was kinetically controlled by
the CO2 diffusion through the sodium crystal phases (Fig. 6).

Fig. 6. CO2 sorption on Na2ZrO3 at different temperatures. (A) T 6 550 C; the


Na2CO3AZrO2 external shell is mesoporous, and the CO2 diffusion occurs through
the mesoporous structure. (B) T > 550 C; the Na2CO3AZrO2 external shell is not
porous [157].

2.2.4. Other ceramic materials


Apart from Li2ZrO3, Li4SiO4, and Na2ZrO3, which are the most
studied ceramic CO2 sorbents up to now, a quite few studies have
been performed on CO2 adsorption by other kinds of ceramic materials including penta-lithium aluminate (Li5AlO4) [158,159],
lithium cuprate (Li2CuO2) [160,161], lithium ferrite (LiFeO2)
[162,163], lithium oxosilicate (Li8SiO6) [164166], lithium orthotitanate (Li4TiO4) [167,168], sodium metatitanate (Na2TiO3) [6,169],
and barium ferrite (Ba2Fe2O5) [170,171]. The CO2 chemisorption on
these materials is similar to that observed for Li2ZrO3, Li4SiO4, and
Na2ZrO3 and leads to the formation of an alkaline carbonate and
the corresponding residual oxide or secondary alkaline phase.
Some of these materials possess different interesting properties
as possible CO2 sorbents. For instance, Li5AlO4 has the best theoretical CO2 capture capacity (0.87 g of CO2/g of sorbent) among the
lithium ceramics because of its high Li/Al molar ratio and the fact
that aluminum is a lighter atom compared to any other element
included in lithium ceramics. Moreover, it is capable of capturing
CO2 in a wide range of temperature (200700 C) [158].
However, Avalos-Rendon et al. [159] reported the loss of CO2 capture capacity during multi-cyclic adsorption/desorption operations
for Li5AlO4. They evaluated the performance of a- and b-Li5AlO4
phases in cyclic adsorption/desorption experiments (adsorption
at 700 C under 100% CO2 and desorption at 750 C under 100%
N2). For a-Li5AlO4 phase, the initial CO2 capture capacity of
47.7 wt.% decreased to 22.1 wt.% after 20 cycles. Under the same
operating conditions, the CO2 capture capacity of b-Li5AlO4 phase
decreased from 62.3 wt.% to 8.1 wt.% over 20 cycles. The reduction
of CO2 capture capacity during the multicycle process was attributed to the sublimation of Li2O during desorption. Li2CuO2, which is
another lithium-based ceramic, has also shown interesting results
as a CO2 sorbent. It has a high theoretical CO2 capture capacity of
0.402 g of CO2/g of sorbent. It should also be mentioned that it is
able to capture CO2 in a wider range of temperatures (120
690 C) in comparison to other lithium-based ceramic materials.
Moreover, it should be noted that copper is lighter and also
cheaper than zirconium. Hence, it may be considered as an option
for industrial applications. However, Palacios-Romero and Pfeiffer
[160] showed that single-phase Li2CuO2 could not be formed by
the co-precipitation method and thus, the prepared sorbent
showed a CO2 uptake capacity much lower than theoretical capture
capacity (0.136 g of CO2/g of sorbent at 650 C). Matsukura et al.

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

[161] applied the solid-state reaction to prepare Li2CuO2 from


Li2CO3 and CuO at 680685 C. The single-phase Li2CuO2 obtained
by this method exhibited CO2 adsorption capacity of 0.402 g of
CO2/g of sorbent, which was equivalent to the theoretical value.
The CO2 capture behavior of Li8SiO6 was studied by
Durn-Muoz et al. [164] for the rst time. They employed a
solid-state reaction method to synthesize a Li8SiO6 sample and
then, evaluated the CO2 capture capacity of developed material.
The results showed that Li8SiO6 might be considered as a new
alternative for CO2 capture at high temperatures because of its
wide temperature range (300700 C), its high CO2 capture capacity (0.521 g of CO2/g of sorbent at 650 C), and its kinetic behavior
(faster adsorption rate in comparison to Li4SiO4). In a further study,
Romero-Ibarra et al. [166] showed that the addition of sodium carbonate, potassium carbonate, or a mixture of both improved the
capture capacity of Li8SiO6 due to the formation of eutectic phases.
Similar to Li8SiO6, Li4TiO4 shows excellent properties for CO2
adsorption. Single phase Li4TiO4 is able to adsorb CO2 in the wide
temperature range of 300850 C, and possesses the high CO2 capture capacity of 0.42 g of CO2/g of sorbent at 856 C. Moreover, its
CO2 adsorption rate is faster than that of Li4SiO4 [167]. However,
the CO2 capture capacity of Li4TiO4 decreases in cyclic operation
because of the decline in the amount of Li4TiO4, which is resulted
from the decrease in the reactable surface area and the vaporization of Li4TiO4, Li2TiO3, and Li2O during multicyclic sorption/desorption processes [168].
Among the ceramics mentioned, the very low CO2 capture
capacity of LiFeO2, Na2TiO3, and Ba2Fe2O5 limits their practical
applications. LiFeO2 readily releases CO2 at lower temperatures
(around 530 C) in comparison with other ceramic materials.
However, its CO2 capture capacity and rate are not sufcient for
practical use. Kato et al. [163] reported the CO2 uptake capacity
of around 0.01 g of CO2/g of sorbent at 500 C for LiFeO2. In a further study, Yanase et al. [162] found that the structural phase transition occurring above 425 C suppressed the CO2 capture of
LiFeO2. Na2TiO3, which is a sodium-based ceramic sorbent, presents lower CO2 capture capacity and slower adsorption/desorption rate in comparison to the most studied sodium-based
ceramic (Na2ZrO3). According to TGA experiments at 600 C,
Na2TiO3 showed the capture capacity of 0.08 g of CO2/g of sorbent,
the sorption rate of 0.1097 wt.%/min, and the desorption rate of
0.043 wt.%/min. However, Na2ZrO3 presented higher adsorption
capacity (0.26 g of CO2/g of sorbent) and faster adsorption
(10.33 wt.%/min) and desorption (1.02 wt.%/min) rates under the
same operating conditions. The low CO2 capture capacity of
Na2TiO3 was ascribed to the presence of high molecular titanates
that are less reactive to CO2 [6]. Ba2Fe2O5 shows CO2 adsorption
in the temperature range of 5001000 C and CO2 desorption
above 1000 C under CO2 partial pressure of 1 atm. Since
Ba2Fe2O5 can capture CO2 at higher temperatures in comparison
to Li4SiO4, a better kinetic behavior is observed for Ba2Fe2O5 under
practical conditions compared to Li4SiO4. However, Ba2Fe2O5
shows lower CO2 capture capacity (0.094 g of CO2/g of sorbent at
1000 C) than Li4SiO4 [170].
2.2.5. Kinetic models
Up to date, different models, including double-shell, multiple
exponential, shrinking core, Avrami-Erofeev, and rate law have
been proposed for analyzing the CO2 adsorption mechanism on
alkaline ceramic materials. As it was mentioned in Section 2.2.1,
Ida and Lin [136] proposed a double-shell model to describe the
mechanism of CO2 adsorption on pure and K-doped Li2ZrO3. Kato
et al. [131] and Essaki et al. [172] used the same model to explain
the CO2 sorption mechanism on Li4SiO4. Double exponential is
another model that has been used by different research groups to
describe the CO2 sorption mechanism on different alkaline

435

ceramics including Li4SiO4, Na2ZrO3, Li8SiO6, and Li5AlO4.


According to this model, two different processes happen during
CO2 capture by alkaline ceramics: (1) CO2 chemisorption over the
surface of ceramics, which produces an external shell containing
alkaline carbonate and a metal oxide or an alkaline secondary
phase; and (2) alkaline element diffusion throughout the external
layer to reach the surface and react with the CO2, which begins
once the external layer is totally formed [9,146,158,165,173
175]. Qi et al. [174] analyzed the reaction mechanism of CO2
adsorption/desorption on Li4SiO4 by comparing the double exponential, shrinking core and AvramiErofeev models. The shrinking
core model, which is commonly used in nonporous materials,
assumes that the rate of reaction is controlled by the rate of chemical reaction. The AvramiErofeev model has been employed for
reactants with highly crystalline structures. According to this
model, the rate of reaction is controlled by the rate of the formation and growth of the reaction product crystals. Fitting the experimental data to the double exponential model showed that the
lithium diffusion process was the limiting step of the whole CO2
sorption process. This observation was inconsistent with the model
assumption that the alkaline diffusion process takes place once the
external shell is completely formed. Shrinking core model did not
t the experimental data well in the whole adsorption temperature
range (550700 C). Hence, the sorption process was not controlled
only by the rate of chemical reaction. The AvramiErofeev model
combined with the double-shell model clearly described the mechanism of the CO2 sorption on Li4SiO4. First, the supercial reaction
between the CO2 molecules and Li4SiO4 led to the formation of
solid Li2CO3 and Li2SiO3 nuclei on the surface. This step was very
short and the rate of CO2 sorption in this step was limited by the
rate of the formation of the product crystals. Then, the diffusion
process took place once the double-shell was formed by the
growth of the Li2SiO3 and Li2CO3 nuclei over the unreacted
Li4SiO4. Pfeiffer and co-workers [164,169,175] have published several papers about the application of the rate law model for analyzing the mechanism of CO2 adsorption on different ceramics such as
Li8SiO6, Na2TiO3, and Na2ZrO3. The rate law model is used when
several processes are involved in the mechanism of CO2
adsorption.

3. Hydrogen production by sorption enhanced steam methane


reforming (SESMR)
The burning of fossil fuels is known to represent the major contributor to the global warming phenomenon. The fossil fuels such
as oil, coal, and natural gas are today the major sources of energy
for industrial activities (over 80% of total industrial energy needs).
However, the drastic reduction of fossil fuels sources and their
harmful inuence on the environment and human health justify
the necessity of alternative energy source development with the
advantages of low emission of pollutants and more energy per
mass. Hydrogen is considered a great candidate for these goals
[176,177]. Since the discovery of this gas in 1766, numerous technological progresses have been made in its production and applications. Recently, the scientic efforts have been concentrated on the
development of new technologies for hydrogen production to
increase the efciency and reduce the production cost [176179].
Hydrogen does not occur free in nature. It can be generally produced by reforming of fuels or non-reforming processes like water
electrolysis and biomass gasication. The fossil fuels (especially
the natural gas) are usually used in industrial applications
[176,177,180182]. Among the three processes for hydrogen production from methane feedstock, including steam reforming (SR),
partial oxidation (POX), and autothermal reforming (ATR), the
most economic process is steam methane reforming (SMR) due

436

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

to the highest thermal efciency and the lowest capital investment


[176,177,180,182]. The main reactions in the steam methane
reforming (SMR) include the endothermic methane reforming (5)
and the exothermic water gas shift (WGS) (6):

Reforming : CH4 g H2 Og $ 3H2 g COg DH 298 206 kJ=mol


5

WGS : COg H2 Og $ H2 g CO2 g DH 298 41 kJ=mol


6
After desulfurization, methane is reformed on nickel-alumina
catalysts at 700900 C and 1420 bar in the presence of steam
(steam to carbon molar ratio (S/C) of 2.5 to 5) to produce syngas,
a mixture of CO and H2. CO can further react with steam to produce
more hydrogen via WGS reaction. The exothermic WGS process is
usually performed in two separate reactors at high (350400 C)
and low (200 C) temperatures, on chromium iron oxide and
copper-zinc catalysts. The efuent gas typically contains 7175%
H2, 47% CH4, 14% CO, and 1520% CO2 (dry basis). For the separation of H2 from CO2, chemical absorption or multicolumn pressure swing adsorption (PSA) are commonly used in the nal
stage of the process, depending on the desired purity [21,177,183].
SMR is a high energy consuming process. Additional steps for
hydrogen purication by chemical absorption (high energy consumption) or PSA (relatively complex process with around 10% loss
of H2) increase the capital investment and reduce the process efciency. The cost of WGS and PSA was estimated to around 30% of
the total cost of H2 production unit [21,184]. An interesting option
to decrease the capital cost of SMR is an integrated process combining the reforming reaction with the in-situ CO2 separation by
high-temperature solid sorbents. The sorption enhanced steam
methane reforming (SESMR) was proposed as a novel efcient
technology to produce highly pure hydrogen by steam methane
reforming [185,186]. The concept of sorption enhanced reaction
process (SERP) is based on the use of a mixture of reforming catalyst and selective regenerable solid sorbent to remove CO2 in-situ
from the reaction zone. The CO2 removal from the gas phase allows
the production of highly pure hydrogen in a single step. The principle of SESMR is to shift the equilibrium of the reversible reactions
(5) and (6) based on the Le Chateliers principle to enhance hydrogen production through in-situ CO2 removal from the reaction
zone, in order to obtain a hydrogen conversion as much as 95%
(dry basis) in a single step, compared to maximum 80% (dry basis)
achieved in a conventional reformer [21]. The selective in-situ CO2
removal from the reaction media can be performed using
high-temperature solid sorbents (MeO represents a metal oxide):

CO2 removal : MeOs CO2 g $ MeCO3 s DH < 0

The simultaneous reactions (5)(7) therefore lead to the overall


reaction (8):

CH4 g 2H2 Og MeOs $ 4H2 g MeCO3 s DH  0


8
The sorbent materials for CO2 separation must be able to resist
in severe operating conditions such as the presence of steam during the adsorption process and high temperature and pressure.
Various high temperature sorbents developed for CO2 capture have
been tested in the SESMR process (the current literature data are
presented in Table 5).
The SESMR process can be performed in dual xed-bed reactors
with periodic switching between hydrogen production and sorbent
regeneration or in parallel circulating uidized-bed reactors [21].
Several important advantages of SESMR process were pointed out
in the literature [21,68,183,185188]:

 highly efcient H2 production with less by-products (CO and


CO2);
 elimination of the individual reactor for WGS;
 elimination of downstream hydrogen purication steps;
 achieving high conversion of methane to hydrogen at signicant
lower temperature (450600 C) compared to traditional SMR
(700900 C);
 replacement of high alloy steels by less expensive materials;
 20% to 25% energy reduction compared to traditional SMR;
 minimization of carbon deposition in the reformer;
 reduction of CO2 release to the atmosphere; relatively pure CO2
can be captured and further sequestrated or used in several
processes;
 reduction of the excess steam used in conventional SMR.
3.1. Application of CaO-based sorbents in SESMR
CaO-based sorbents are the most well-known CO2 sorbents used
for the H2 production by SESMR process so far [187,189191]. In an
early study, Balasubramanian et al. [192] used CaO sorbent obtained
from the calcination of high purity CaCO3. At 650 C and S/C ratio of 4
the authors reported the production of H2 with a molar fraction
higher than 95% (dry basis). The inuence of several operating
parameters (temperature, steam to carbon ratio, and feedstock composition) was investigated in order to nd the optimum working
conditions. However, as it was mentioned before, the CO2 capture
capacity of CaO-based sorbents decreases in cyclic operation, particularly under severe regeneration conditions (the presence of CO2 in
the regeneration atmosphere) because of the sintering of the CaO
particles and thus, frequent shut down is needed for regenerating
the CO2 sorbent. Therefore, many attempts have been made to
develop thermal-stable CaO-based sorbents for application in
SESMR process. Li et al. [193] studied the application of
CaO/Ca12Al14O33 (75%/25%) sorbent developed previously [12,65]
in a continuous SESMR process for a period of 400 min. The process
was cyclically operated in two parallel xed-bed reactors: (i) hydrogen production and CO2 sorption at 630 C, 1 atm, and S/C ratio of 5
and (ii) sorbent regeneration under argon at 850 C and 1 atm. The
authors reported that hydrogen with a purity of 95% could be continuously produced. It was concluded that the switchover time
(pre-breakthrough period) between the two reactors was a key
parameter for H2 efciency. Broda et al. [194] evaluated the performance of a mixture containing 5.7 g of Ni-hydrotalcite-derived catalyst and 1.26 g of Ca-based sorbents in SESMR process for 10
successive cycles. The Ni-hydrotalcite (Ni-Htlc)-derived catalyst
containing 47 wt.% of Ni was developed by co-precipitation method
from Ni(NO3)26H2O, Mg(NO3)26H2O, and Al(NO3)39H2O precursors. Al2O3-stabilized Ca-based sorbents with the Ca2+/Al3+ ratio of
90:10 or 80:20 were prepared via solgel technique from aluminum
isopropoxide and calcium hydroxide precursors. Under SESMR conditions (550 C, S/C of 4), H2 was produced with a purity of 99% (dry
basis). The H2 production rate declined by 1.9% per each cycle when
the mixture of Ni-Htlc and Ca:Al 80:20 was used, representing a
decrease of 275% in comparison with Ni-Htlc/limestone mixture.
The good activity of the Al-stabilized CaO-based sorbents under
SESMR conditions was on account of the homogeneously dispersed
high Tammann temperature support (Ca12Al14O33), which hindered
the structural sintering and pore pluggage and stabilized its nanostructured morphology. Xu et al. [195] developed a series of
Ca9Al6O18-CaO sorbents from different calcium precursors, including calcium acetate, calcium citrate, calcium lactate, and calcium
gluconate by a solgel method. Among all developed sorbents, the
calcium lactate-derived sorbent showed the best CO2 uptake performance in comparison with the other sorbents tested because of its
largest surface area and pore volume. A mixture of 3 g of calcium
lactate-derived sorbent containing 90 wt.% CaO and 1 g of

437

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444


Table 5
Summary of investigations on SESMR.
Patterna NiO loading
(wt.%)

Reaction conditions

Ni commercial cat./com. CaO


Ni commercial cat./dolomite

M
M

22.0
18

Ni commercial cat./CaO:Ca12Al14O33
Ni commercial cat./dolomite
Ni commercial cat./CaO:Ca12Al14O33
NiO/CaO/Ca12Al14O33
NiO/CaO
La2O3/NiO/CaO/Al2O3
ZrO2/NiO/CaO/Al2O3
NiO/CaO/HTlc (Al-Mg)
NiACaOACa12Al14O33
NiOAHTlc (AlAMg)/Al-stabilized CaO
NiOAHTlc (AlAMg)/CaO:Ca9Al6O18
Ni commercial cat./Li2ZrO3
Ni commercial cat./Na2ZrO3
NiOAHTlc (AlAMg)/Li2ZrO3
NiOAHTlc (AlAMg)/Na2ZrO3
NiO-c-Al2O3/CaOACa9Al6O18
NiAMgAlO/CaOACa9Al6O18
NiAHtlc/CaOACa12All4O33

M
M
M
H
H
H
H
H
H
M
M
M
M
M
M
M
M
M

450750 C, 15 atm., S/C: 35 N.A.


650 C; 15 atm., S/C: 4
800950 C, 100% N2,
4% O2/N2, 100% CO2
630 C; 1 atm., S/C: 5
850 C, 100% Ar
600 C, 1 atm., S/C: 3
850 C, 100% N2
650 C, 1 atm, S/C: 3.4
850 C, 100% He
650 C, 1 atm, S/C: 3.4
850 C, 100% He
600 C, 1 atm, S/C: 3
N.a.
600 C, 1 atm, S/C: 4
800 C, 100% N2
600 C, 1 atm, S/C: 4
800 C, 100% N2
550 C, 1 atm, S/C: 4
750 C, 100% N2
630 C, S/C: 3
780 C, 100% N2
550 C, 1 atm, S/C: 4
750 C, 100% N2
550 C, 1 atm, S/C: 4.2
800 C, 20% H2/80% N2
505 C, 1 atm, S/C: 4
N.a.
600 C, 1 atm, S/C: 4
N.a.
575 C, 5 atm, S/C: 5
650 C, 100% Ar
575 C, 5 atm, S/C: 5
750 C, 100% Ar
650 C, 5 atm, S/C: 2.85
850 C, 100% N2
600 C, 1 atm, S/C: 4
800 C, 20% H2/80% N2
550 C, S/C: 4
750 C, 100% N2

NiOACaO/Ca9Al6O18
CaOAZr/NiO

H
H

20
(a)
18
20
(b)
N.a
(c)
(d)
7
(e)
(g)
N.a.
N.a.
(f)
(f)
5
N.a.
47 wt.% in
catalyst (for Ni)
25
650 C, 1 bar, S/C: 4
20.5
650 C, 1 bar, S/C: 4

Solid (sorbent, catalyst)

Reac.

Reg.

Ref.
Reactorb Number H2
of cycles concentration
(molar %)
FB
FB

1
25

>95
>95

[192]
[190]

FB
FB
FB
FB
FB
FB
FB
FB
FB
FB
FB
FB
FB
FB
FB
FB
FB
PB

12
4
13
1
1
30
20
10
4
10
1
1
1
1
1
1
35
10

>90
>98
>92
90
80
>92
>90
99
95
99
97
85
97
68
97
89.1
98
99

[193]
[187]
[191]
[201]
[199]
[202]
[56]
[203]
[74]
[196]
[200]
[207]
[207]
[128]
[128]
[206]
[195]
[194]

800 C, 11.1% H2 (in Ar) FB


800 C, 11.1% H2 (in Ar) FB

30
10

>95
>95

[204]
[205]

(a) Ni loading: >12 wt%; (b) Ni loading: 12.5 wt%; (C) Ni loading: 15 wt%; (d) Ni loading: 45 wt%; (e) Ni loading: 47 wt%; (f) Ni loading: 40 wt%; (g) Ni:Mg:Al (atomic ratio in
HTlc catalyst) = 0.5:2.5:1
a
Mixture (M) or hybrid (H).
b
Fixed-bed (FB) or Packed-bed (PB).

Ni-based catalyst was tested in the SESMR process at 600 C and S/C
of 4. The H2 concentration during the prebreakthrough stage was
98%. During 10 successive reforming-regeneration cycles, the
methane conversion and different species concentrations remained
almost constant, demonstrating the good performance of the
CaO-based sorbent developed by solgel technique. Broda et al.
[196] used a mixture of Ni-HTlc-derived catalyst (47 wt.% Ni, prepared by co-precipitation) and a synthetic Ca-based sorbent (CaO
supported on calcium aluminate). The sorbent pellets were synthesized from limestone and commercial calcium aluminate (CA-14:
71% Al2O3 and 28% CaO). At 550 C and S/C of 4, 99 vol.% H2 (dry
and N2-free basis) was obtained in the reforming process. After 10
sorption/regeneration cycles, the CO2 capture capacity was 0.41 g
of CO2/g of sorbent, in comparison with limestone (0.22 g of CO2/g
of sorbent). The appropriate thermal stability of CO2 sorbent was
attributed to the uniform dispersion of Ca12Al14O33 among CaO
particles.
For large-scale hydrogen production units where mass transfer
limitations can signicantly affect the process efciency, the catalystsorbent mixing conguration is an important parameter to
be considered. Besides a physical admixture of catalyst with sorbent, hybrid catalystsorbent patterns which integrate the catalytic
reaction and the CO2 sorption in a single pellet can present some
advantages such as the elimination of mass diffusional limitation
and the reduction of reactor volume [197,198]. Several studies have
been performed on the development of hybrid catalystsorbent
materials for SESMR process [199206]. A rst attempt to combine
CaO-based sorbent and catalyst in a single pellet was performed by
Martavaltzi and Lemonidou [201] who developed a hybrid
NiACa12Al14O33ACaO catalystsorbent. The optimum NiO loading
of 20 wt.% was shown to lead to a H2 concentration of 90% at
650 C and S/C of 3.4, as well as CO2 and CO efuent concentrations
of 2.8% and 2%, respectively. The process was only limited to one
cycle. Wu and Wang [56] further investigated the application of
ZrO2-stabilized NiOACaO/Al2O3 hybrid catalystsorbent in the
SESMR during 20 cycles. The authors reported a H2 concentration

of more than 90%. The favorable stability and activity of the proposed hybrid material during cyclic operation was found to be on
account of avoiding the formation of NiAl2O4 phase by incorporating ZrO2 particles. Feng et al. [202] developed a La2O3-stabilized
NiO-CaO/Al2O3 hybrid catalystsorbent prepared by two-step
impregnation of lanthanum and nickel precursors. At 600 C and
S/C of 4, H2 concentration of more than 92% was achieved during
30 SESMR cycles. The incorporation of La2O3 could improve both
stability and nickel grain dispersion over the substrate.
Besides the suitability of CO2 sorbent, appropriate Ni dispersion
and surface area of catalyst are also important parameters in H2
production. A HTlc-based hybrid catalystsorbent was synthesized
by Broda et al. [203] via co-precipitation technique. The hybrid
material contained both Ni reforming catalyst and Ca-based CO2
sorbent dispersed in the HTlc structure containing Mg and Al
(Fig. 7). The appropriate surface area (54 m2/g) and the high dispersion of Ni and Ca in the HTlc structure resulted in a high H2 production efciency of 99 vol.% (a dry basis composition is always
mentioned, if otherwise specied) and adequate thermal stability
over cyclic SESMR operation. The proposed hybrid material produced a better H2 purity compared to a mixture of limestone and
NiSiO2 or nickel HTlc-derived catalyst. However, the loading of
CaO in the hybrid structure was only 21 wt.%, requiring the use of
a high amount of hybrid material in the reaction to obtain an adequate CO2 adsorption capacity. As stated by the authors, the CO2
uptake capacity of the hybrid material averaged only 0.074 g of
CO2/g of sorbent over 10 cycles. Kim et al. [74] studied the synthesis
of hybrid CaOACa12Al14O33ANi composite from calcium nitrate
tetrahydrate, aluminum nitrate nonahydrate, and Ni precursor by
combination of precipitation and hydration methods, and its application in the SESMR (S/C of 3, 630 C). Ca12Al14O33 made spacious
pathway available for CO2 diffusion via forming porous structure,
thus providing an excellent cyclic stability for Ca-based sorbent.
The SESMR experiments using hybrid CaOACa12Al14O33ANi composites with different loading of Ni precursor (3, 5, 7, and
10 wt.%) revealed that 7 wt.% of Ni loading led to the best

438

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

Fig. 7. Schematic diagram of hybrid catalyst-CO2 sorbent arrangement [203].

performance (CH4 conversion and H2 production). The H2 concentration was maintained at 9495% for 70 min (prebreakthrough)
during SESMR. It was also found that the presence of a high Ni loading (10 wt.%) caused Ni self-agglomeration. Radfarnia and Iliuta
[204] developed NiOACaO/Ca9Al6O18 hybrid catalystsorbent
materials with various nickel loading (12, 18, and 25 wt.% NiO)
via a simple wet-mixing method (limestone acidication coupled
with two-step calcination technique) and investigated their application in SESMR operation (S/C of 4, 650 C and 1 bar). According
to the experimental results, the best performance belonged to the
hybrid material containing 25 wt.% NiO, which showed an average
H2 production efciency of 96.1% during 10 cycles (180 min reaction per cycle). Moreover, the long-term application (30 cycles,
45 min reaction per cycle) led to an average H2 production efciency of 97.3%, proving its high efciency in the SESMR process.
The excellent performance of this catalystsorbent hybrid material
was attributed to the easy access of Ni sites resulted from higher Ni
loading and proper distribution of Ni on the substrate. The same
authors also developed a CaOAZr/Ni (13, 18, 20.5 wt.% NiO) sorbentcatalyst material by wet-mixing/sonication technique and
its application in the SESMR showed that the one with 20.5 wt.%
NiO loading presented the most suitable activity (H2 yield of 91%
at the end of the 10th cycle) [205].
In a recent study, Barelli et al. [206] employed wet mixing methods to prepare three different catalystsorbent materials for
SESMR:
NiOACaO/Ca9Al6O18
composite
(M1),
NiOACaO/
Ca12Al14O33 composite (M2), and a physical mixture of
CaO/Ca9Al6O18 and NiO/cAAl2O3 (M3). M3 showed the best performance in cyclic carbonation/decarbonation experiments, where the
initial CO2 capture capacity of 0.55 g of CO2/g of sorbent decreased
to 0.495 g of CO2/g of sorbent after 14 cycles (reduction of 17%). For
comparison, the CO2 capture capacity of standard pure CaO
decreased from 0.49 g of CO2/g of sorbent to 0.288 g of CO2/g of sorbent over 14 cycles (reduction of 57.5%). XRD, SEM and TEM analysis of the used materials justied the performance of all three
materials during cyclic CO2 adsorption/desorption process. While
no change was observed in the porous structure and morphology
of M3 sorbent, agglomeration of CaCO3 on the surface of
Ca9Al6O18 and partial loss of roughness and porosity of used M1
led to much lower cyclic CO2 adsorption/desorption performance
in comparison with M3. In addition, the formation of NiAl2O4 and
the deterioration of the porous structure during the cyclic CO2
adsorption/desorption process caused the poorest performance of
M2. The material with the best performance (M3) was further
applied in SESMR process at different temperatures between 500
and 650 C. Although the molar fraction of H2 at the reactor outlet
reached 89.1% during the pre-breakthrough period at 650 C (S/C
of 2.85), it was still much lower than that at equilibrium. This could
be due to difcult contact between gas and solid because of the particle size of the catalyst, low catalyst load and residence time.

3.2. Application of ceramic sorbents in SESMR


Compared to CaO-based materials, only few experimental data
concerning the application of ceramic sorbents in the SESMR are
available in the literature. The application of different
alkaline-containing sorbents was studied by Yi et al. [207] in a
xed-bed reactor. The use of Na2ZrO3 prepared by liquidliquid
synthesis method resulted in a H2 concentration of 96.8% at
600 C, 1 bar and S/C of 4. However, data concerning the cyclic
adsorption/regeneration were not reported. It was found that during regeneration at high temperatures Na migrated and covered
the active Ni-sites, which resulted in the catalyst deactivation.
The application of Li2ZrO3 prepared by the same synthesis method
revealed the inadequacy of this sorbent due to the very low CO2
capture kinetics in the operating conditions (low partial pressure
of CO2). An initial H2 concentration of less than 85% (tending
towards the SMR equilibrium conversion of 73%) indicated that
CO2 could not be efciently removed from the reaction zone while
it was generated in the reforming process. Another research group
from the Norwegian University of Science and Technology (NTNU)
also showed interest in the synthesis and application of
alkaline-based sorbents in the SESMR. Ochoa-Fernandez et al.
[137] performed simulations of SESMR using CO2 sorption data
obtained experimentally on different kinds of materials like
Li2ZrO3, K-Li2ZrO3, Na2ZrO3, Li4SiO4 and CaO. The favorable thermodynamic behavior of sorbents at low CO2 partial pressure was
found to play a key role in the success of the SESMR process.
Although the use of CaO resulted in the highest H2 yield (above
98%) at 575 C, 10 bar and S/C ratio of 5, its signicant uptake
capacity reduction would be a serious problem in cyclic operation.
For comparison, the simulation results obtained for the other considered sorbents were: K-Li2ZrO3 (93%), Na2ZrO3 (90%), Li2ZrO3
(89%), and Li4SiO4 (82%). Based on CO2 sorption data, it was concluded that Na2ZrO3 might be an alternative to CaO due to the good
kinetic behavior and the appropriate stability during sorption/desorption cycles. However, the same group concluded later that the
stability of Na2ZrO3 was signicantly reduced during the operation
at high steam pressure [208]; this was however not in agreement
with data reported by Yi et al. [134]. More thorough investigations
on the effect of the steam on ceramic sorbents during cyclic SESMR
operation would be necessary to clarify their behavior. The application of Na2ZrO3 and Li2ZrO3 sorbents prepared by liquidliquid
synthesis was also studied by Ochoa-Fernandez et al. [128,137].
The use of Na2ZrO3 at 575 C, 5 bar, and S/C of 5, with sorbent
regeneration at 750 C under Ar, resulted in a H2 yield above
97%. For Li2ZrO3, the H2 yield was very close to the thermodynamic
equilibrium of traditional SMR process at the reaction conditions
investigated (67.9%), due to the kinetic limitation during the CO2
sorption, in agreement with the results of Yi et al. [134]. It is
important to note that a drastic decrease of H2 yield was observed

439

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

during the second cycle of operation for both Na2ZrO3 and Li2ZrO3.
Even if the reason was not very clear, this behavior seemed to be
due to catalyst poisoning in the presence of impurities present in
the sorbents.
4. Hydrogen production by sorption enhanced steam glycerol
reforming (SESGR)
As mentioned in the previous section, methane reforming is
commonly used in industrial applications for hydrogen production
[176,177,180182]. The steam methane reforming provides about
48% of worldwide produced hydrogen [209]. However, hydrogen
production from methane deals with serious difculties such as
environmental problems and drastic reduction of fossil fuel
sources. Therefore, extensive works were directed on the development of alternative renewable sources such as water and biomass
or biomass-derived oxygenates. Glycerol, a by-product of the transesterication of renewable biological sources (i.e., vegetable oils
and animal fat oils), may be considered an interesting renewable
source of hydrogen [210214]. The production of 10 kg of biodiesel
generates around 1 kg of crude glycerol [210]. Hence, the rapid
growth in the production of biodiesel from 2000 year has resulted
in great increase of crude glycerol. It is anticipated that the annual
production of glycerol will reach to about 3 megatons in 2020,
whereas the industries consume only 500 kilotons glycerol each
year [214,215]. The increase of the availability of this industrial
waste, as well as the renewability and the low cost of glycerol, make
it very attractive as an alternative source of hydrogen [211].
Pyrolysis, partial oxidation, steam reforming, autothermal
reforming, and aqueous-phase reforming are some promising
methods for converting glycerol into hydrogen. By now, steam
reforming is the most common method for converting glycerol into
hydrogen [213,215,216]. This global process consists of complex
reactions, which lead to the formation of several by-products and
therefore, affect the nal purity of H2 adversely [213]. Glycerol
pyrolysis (9) and WGS (6) are the main reactions in this process:

C3 H8 O3 $ 4H2 3 CO DHo298 251 kJ=mol


CO H2 O $ H2 CO2

DH 298 41 kJ=mol

9
6

Consequently, the overall reaction of glycerol reforming is represented by Eq. (10), where one mole of glycerol can theoretically
produce 7 mol of hydrogen:

C3 H8 O3 3H2 O $ 7H2 3 CO2

DH 298 128 kJ=mol

10

The side reactions that can occur in the steam reforming of glycerol are presented in Table 6. The steam glycerol reforming process
is carried out at temperatures of 500900 C, 1 atm and
water/glycerol molar ratio of 69, in the presence of Ni, Co and
noble metals (such as Pt, Pd and Rh) based catalysts. H2, CO, CO2,
and CH4 are the main gaseous products [213216].
One of the obstacles to the utilization of hydrogen obtained from
steam glycerol reforming for energy production is the high CO and
CO2 content. In particular, the presence of large amount of CO2 signicantly drops the efciency of fuel cells, while CO strongly poisons the catalyst of proton-exchange membrane fuel cells
(PEMEC) [217]. Furthermore, the cost of hydrogen separation from
a H2-rich gas containing impurities causes major cost penalties.
Therefore, a better system is needed to achieve high purity hydrogen. The application of the sorption enhanced process is therefore
an interesting option to produce high purity hydrogen in a single
step [210,211,214]. Compared to the traditional steam reforming,
the in-situ CO2 removal increases glycerol and steam conversions
as well as hydrogen purity [214,215]. Moreover, the in-situ CO2 capture during the steam glycerol reforming decreases the risk of coke

Table 6
Side reactions during steam glycerol reforming.
Entry

Reaction

DH298 (kJ/mol)

1
2
3
4
5
6
7

CO 3H2 $ CH4 H2 O
CO2 4H2 $ CH4 2H2
H4 CO2 $ 2CO 2H2
2CO $ CO2 Cs
CH4 $ 2H2 Cs
CO H2 $ H2 O Cs
CO2 2H2 $ 2H2 O Cs

206
165
247
172
75
131
306

formation and the reforming reactions can be carried out at relatively low steam/carbon ratios [212]. Dou et al. [214] provided an
overview on some issues and challenges of SESGR process such as
selecting suitable sorbents, extending operation time, and nding
a way for continuous reaction/regeneration in order to achieve
high-efciency hydrogen production from SESGR process.
To date, much fewer studies have been performed on the
sorption enhanced steam glycerol reforming (SESGR) compared to
SESMR. In the presence of CaO as the CO2 sorbent, it was found that
the in-situ CO2 removal resulted in a signicant enhancement of H2
production and thermal efciency as well as an important reduction of CO concentration [210,218220]. Chen et al. [221] performed thermodynamic analyses on the SESGR process based on
the principle of Gibbs free energy minimization for chemical reactions. They evaluated the effect of temperature (327727 C), pressure (14 bar), S/C ratio (14), percentage of CO2 removal through
adsorption (099%), and carrier gas to feed reactants molar ratio
(15) on the reforming reactions and carbon formation. The results
demonstrated that the in-situ CO2 removal led to the enhancement
of glycerol conversion to hydrogen. The maximum hydrogen yield
was increased from 6 to 7 moles/mole of glycerol by the in-situ CO2
removal via adsorption. The analyses proposed that the most
favorable conditions for the SESGR process were: temperature
range around 527577 C at atmospheric pressure and S/C of
around 3.0. The most favorable temperature for the SESGR process
was approximately 100 C lower than that for the traditional steam
glycerol reforming without in-situ CO2 removal. As carbon formation can occur at low S/C ratios, the in-situ CO2 removal can considerably reduce the lower limit of the S/C ratio to limit the carbon
formation [221]. He et al. [211] studied the SESR process of pure
glycerol by using the CoANi/HTls as reforming catalysts and calcined dolomite as CO2 sorbent, at 500650 C and S/G of 3, 4, and
9. The CoANi/HTls catalysts (25%Co15%Ni/HTls and 30%Co
10%Ni/HTls)
were
developed
by
co-precipitation
from
Ni(NO3)26H2O, Co(NO3)3, Mg(NO3)36H2O and Al(NO3)39H2O.
According to the experimental results, the hydrogen purity reached
approximately the theoretical value at temperatures beyond
575 C or S/G no less than 4 (for example, both hydrogen purity
and yield reached around 99% for a S/G of 9). Such a high H2 yield
obtained in the SESGR process was attributed to the enhancement
of methane steam reforming and watergas shift reactions and the
elimination of non-catalytic reactions in the SESR process. Below
550 C and a low S/G of 3, hydrogen production was compromised
because of the high tendency of pyrolysis and the low efciency of
steam reforming. Dou et al. [222] studied the SESGR process in a
xed-bed reactor using a mixture of commercial Ni-based catalyst
(18 wt.% NiO/82 wt.% Al2O3) and CaO-based sorbent derived from
dolomite (weight ratio of catalyst to sorbent of 1:1), at temperatures in the range of 400700 C, 1 atm, and S/C of 3. The experimental results demonstrated that the SESGR is an efcient
process for obtaining hydrogen purity of more than 90%. 500 C
was shown to be the optimum temperature, with the highest H2
purity of 97% and the longest CO2 breakthrough period.
One of the most important parameters that affect the economic
benets of H2 production by SESGR is the price of glycerol. Up to

440

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

date, the market price of crude glycerol is between 25 and 46 $/ton,


which is signicantly lower than that of puried glycerol (55
78 $/ton) [212]. The SESR of crude glycerol becomes therefore
more attractive for a low price H2 production. Fermoso et al.
[223] and Dou et al. [224] showed the great potential of crude glycerol for producing H2 with high yield and purity. Fermoso et al.
[223] evaluated the SESR of crude glycerol in a xed-bed reactor
over a mixture of Ni-Co/HTlc catalyst and calcined dolomite as
CO2 sorbent (sorbent/catalyst = 5 g/g), and at temperatures
between 550 and 600 C, 1 atm, and S/C of 3. The experimental
results revealed that the in-situ CO2 removal through adsorption
signicantly enhanced the reforming and watergas shift reactions, allowing high H2 yield (88%) and purity (99.7 vol%). In a similar work, Dou et al. [224] evaluated the steam reforming of crude
glycerol in a xed-bed reactor over a commercial Ni-based catalyst
(18 wt.% NiO/82 wt.% Al2O3) at a temperature range of 400700 C,
1 atm, and S/C of 3, with and without in-situ CO2 removal.
CaO-based sorbent derived from dolomite was used for the
SESGR. In the absence of CO2 removal, the H2 purity and the crude
glycerol and steam conversions were 68%, 100% and 11% at 600 C,
respectively. The in-situ CO2 removal during the steam reforming
led to a hydrogen purity beyond 88% in pre-breakthrough conditions. It is important to mention that, under the same SESGR conditions, the hydrogen yield and purity obtained from crude glycerol
were lower compared to pure glycerol [211,222]. This was attributed to the increased coke formation resulted from the high coking
potential of heavy components present in the crude glycerol (such
as fatty acid methyl esters). The production of almost pure H2 from
crude glycerol is therefore a great challenge because of the complexity of crude glycerol composition.
The previous works have investigated H2 production by SESGR
in a xed-bed reactor using a mixture of catalyst and CO2 sorbent.
The operating time for high purity H2 production is quite short
because of the limited capacity of the CO2 sorbent. In a continuous
SESGR process, which is based on the concept of a continuous ow
of catalyst and sorbent for both reaction and regeneration using two
moving bed reactors, the high purity H2 production is expected to
be extended for a longer period of time. Dou et al. [225] investigated
the continuous SESGR process using a mixture of Ni-based catalyst
(NiO/NiAl2O4) and CaO-based sorbent (catalyst to sorbent weight
ratio: 1:1). The Ni-based catalyst containing 42.1 wt.% NiO was
developed by co-precipitation from Ni(NO3)26H2O and
Al(NO3)39H2O precursors. The tests were carried out at 500 and
600 C, with S/C of 3. The simultaneous regeneration of catalyst
and sorbent was performed at 900 C with the gas mixture of N2
and steam. In the SESGR process performed in a xed-bed reactor,
in the pre-breakthrough time that lasted for only 10 min the
amounts of CO2, CO and CH4 decreased and H2 concentration was
considerably increased in comparison to conventional SGR.
However, in the continuous SESGR, the mixture of catalyst and sorbent was moved constantly between the reforming and regeneration moving-bed reactors and the H2 production was not
interrupted for the regeneration of both catalyst and sorbent.
During the operating time of 60 min, hydrogen purity was 93.9%
and 96.1% at 500 and 600 C, respectively. However, the maximum
conversion of CaO sorbent was only 15.5% due to the very short residence time of the sorbent in the reformer. The continuous carbonation and calcination of CaO at this low-level conversion did not
exhibit a signicant drop of the sorbent reactivity. In a further
study, Dou et al. [226] proposed a new continuous system, the
enhanced sorption chemical looping steam reforming process. The
mixture of catalyst and sorbent (weight ratio of 1:1) were moved
continuously between two moving-bed rectors. In the reforming
reactor, oxidation, steam reforming, watergas shift and in-situ
CO2 removal were combined and carried out at 500600 C and
S/C of 1.53.0. NiO/NiAl2O4 was used as catalyst and limestone as

sorbent. In the air reactor, sorbent regeneration, catalyst oxidization and coke combustion were carried out at 900 C in an air atmosphere. The NiO/NiAl2O4 catalyst containing 42.1 wt.% NiO was
developed by co-precipitation from Ni(NO3)26H2O and
Al(NO3)39H2O precursors. The experimental results showed that
the increase of temperature and S/C value enhanced the H2 purity.
The best results (hydrogen purity higher than 90%) were obtained
at temperatures of 500600 C and S/C of 1.53.0.

5. Conclusion and recommendations for future works


The sorption enhanced steam reforming is an integrated process
combining reforming, water gas shift and CO2 capture in order to
produce highly pure H2 in a single step. These reactions occur
simultaneously over a mixture of reforming catalyst and CO2 sorbent. This hybrid process has several important advantages in
comparison with the traditional steam reforming, especially the
energy efciency improvement and the reduction in capital cost.
Hydrogen can be obtained with a purity of 9598% (dry basis) with
very low carbon oxides content, compared to maximum 80% (dry
basis) for traditional steam reforming. Further purication is usually not required in most applications. Moreover, H2 production
is accompanied by the generation of a rather pure CO2 stream suitable for storage and further use.
Several CO2 sorbents have been studied as potential candidates,
including Ca-based oxides and mixed alkaline oxides of Li and Na.
Ca-based sorbents derived from natural precursors are especially
advantageous due to the low cost and availability. Despite the high
CO2 capacity and good kinetics over a wide range of temperatures
and pressures for Ca-based sorbents, they suffer from the important
drawback of high instability in long-term operation due to the sintering phenomena at high temperatures. In this context, a lot of
works have been directed on the improvement of their durability
in cyclic operation by using various calcium precursors, incorporation of stable inert materials into CaO structure, and using different
kinds of treatments (like steam hydration, thermal pretreatment,
recarbonation and treatment with chemical solutions). However,
the review of available data showed some contradictory results
about the effect of treatments such as steam hydration during calcination. Further systematic studies are therefore highly needed to
conrm or inrm specic behaviors. Moreover, as most investigations related to this kind of sorbents were performed in operating
conditions far from industrial applications (limited reaction time
and low CO2 partial pressures), extended works have to be done
on wider range of operating parameters.
Alkaline ceramics represent another group of high-temperature
CO2 sorbents that have been attracted interest especially because
of their superior stability in cyclic operation compared to CaO.
Li2ZrO3, Li4SiO4, and Na2ZrO3 are the most studied ceramics materials. However, most of them suffer from slow kinetics especially at
low CO2 partial pressures. Compared to Ca-based sorbents, very little information is available concerning their application in the
sorption enhanced processes. More thorough investigations on
the effect of the steam on ceramic sorbents during cyclic operation
would especially be necessary to clarify their behavior.
Based on the available works related to the use of CO2 sorbents
in the SESMR process, it has been demonstrated that nearly pure
H2 product can be potentially obtained. Simple mixing patterns
of sorbent and catalyst are used in most investigations. However,
in order to eliminate the mass transfer limitations, hybrid catalystsorbent structures have recently appeared as an interesting
option with the benets of decreasing both reactor volume and
cost of operation. The development of this kind of materials is still
in embryonic stage. Exhaustive studies are highly needed in the
development of efcient hybrid catalystsorbent materials and

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

the evaluation of their efciency and stability under severe operation conditions and cyclic operation.
In terms of the application of CO2 sorbents in the SESGR process,
few works have been done by using CaO-based sorbents and various catalysts. In all these works, the catalyst and sorbent were
mixed physically and none of them considered using hybrid catalystsorbent arrangement up to now. In this context, systematic
studies on the development of efcient hybrid catalystsorbent
materials and the evaluation of their stability under severe operating conditions and cyclic operation represent interesting opportunities for research. Although the SESR of crude glycerol is more
attractive compared to pure glycerol to obtain H2 with a lower
price, the SESR of crude glycerol has been very scarcely investigated. The main challenge in the SESR of crude glycerol is the carbon formation because of the presence of heavy compounds, such
as fatty acid methyl esters. Therefore, the production of pure H2
from crude glycerol is still a challenge due to the composition complexity of crude glycerol.
Acknowledgments
Financial support from Natural Sciences and Engineering
Research Council of Canada (NSERC), FRQNT Centre in Green
Chemistry and Catalysis (CGCC) and Centre de Recherche en
Catalyse et Chimie Verte (C3V, Laval University) is gratefully
acknowledged.
References
[1] Global greenhouse gas reference network. <http://www.esrl.noaa.gov/gmd/
ccgg/trends/weekly.html>, 2015.
[2] S.I. Plasynski, J.T. Litynski, H.G. McIlvried, R.D. Srivastava, Progress and new
developments in carbon capture and storage, Crit. Rev. Plant Sci. 28 (2009)
123138.
[3] B. Metz, O. Davidson, H. De Coninck, M. Loos, L. Meyer (Eds.), IPCC Special
Report on Carbon Dioxide Capture and Storage, Cambridge University Press,
UK, 2005.
[4] K. Nakagawa, T. Ohashi, A novel method of CO2 capture from high
temperature gases, J. Electrochem. Soc. 145 (1998) 13441346.
[5] E. Ochoa-Fernandez, M. Ronning, T. Grande, D. Chen, Synthesis and CO2
capture properties of nanocrystalline lithium zirconate, Chem. Mater. 18
(2006) 60376046.
[6] A. Lopez-Ortiz, N.G.P. Rivera, A.R. Rojas, D.L. Gutierrez, Novel carbon dioxide
solid acceptors using sodium containing oxides, Sep. Sci. Technol. 39 (2004)
35593572.
[7] T.J. Zhao, E. Ochoa-Fernandez, M. Ronning, D. Chen, Preparation and hightemperature CO2 capture properties of nanocrystalline Na2ZrO3, Chem. Mater.
19 (2007) 32943301.
[8] M. Kato, K. Nakagawa, New series of lithium containing complex oxides,
lithium silicates, for application as a high temperature CO2 absorbent, J.
Ceram. Soc. Jpn. 109 (2001) 911914.
[9] R. Rodriguez-Mosqueda, H. Pfeiffer, Thermokinetic analysis of the CO2
chemisorption on Li4SiO4 by using different gas ow rates and particle
sizes, J. Phys. Chem. A 114 (2010) 45354541.
[10] M. Olivares-Marin, T.C. Drage, M.M. Maroto-Valer, Novel lithium-based
sorbents from y ashes for CO2 capture at high temperatures, Int. J.
Greenhouse Gas Control 4 (2010) 623629.
[11] A. Silaban, M. Narcida, D.P. Harrison, Characteristics of the reversible reaction
between CO2(g) and calcined dolomite, Chem. Eng. Commun. 146 (1996)
149162.
[12] Z.S. Li, N.S. Cai, Y.Y. Huang, H.J. Han, Synthesis, experimental studies, and
analysis of a new calcium-based carbon dioxide absorbent, Energy Fuels 19
(2005) 14471452.
[13] H. Lu, E.P. Reddy, P.G. Smirniotis, Calcium oxide based sorbents for capture of
carbon dioxide at high temperatures, Ind. Eng. Chem. Res. 45 (2006) 3944
3949.
[14] J.M. Valverde, Ca-based synthetic materials with enhanced CO2 capture
efciency, J. Mater. Chem. A 1 (2013) 447468.
[15] A.M. Kierzkowska, R. Pacciani, C.R. Muller, CaO-based CO2 sorbents: from
fundamentals to the development of new, highly effective materials,
ChemSusChem 6 (2013) 11301148.
[16] F.C. Yu, N. Phalak, Z.C. Sun, L.S. Fan, Activation strategies for calcium-based
sorbents for CO2 capture: a perspective, Ind. Eng. Chem. Res. 51 (2012) 2133
2142.
[17] W.Q. Liu, H. An, C.L. Qin, J.J. Yin, G.X. Wang, B. Feng, M.H. Xu, Performance
enhancement of calcium oxide sorbents for cyclic CO2 capture a review,
Energy Fuels 26 (2012) 27512767.

441

[18] S. Wang, C. An, Q.-H. Zhang, Syntheses and structures of lithium zirconates for
high-temperature CO2 absorption, J. Mater. Chem. A 1 (2013) 35403550.
[19] L. Barelli, G. Bidini, F. Gallorini, S. Servili, Hydrogen production through
sorption-enhanced steam methane reforming and membrane technology: a
review, Energy 33 (2008) 554570.
[20] N.D. Hutson, B.C. Attwood, High temperature adsorption of CO2 on various
hydrotalcite-like compounds, Adsorption J. Int. Adsorpt. Soc. 14 (2008) 781789.
[21] D.P. Harrison, Sorption-enhanced hydrogen production: a review, Ind. Eng.
Chem. Res. 47 (2008) 64866501.
[22] J.C. Abanades, The maximum capture efciency of CO2 using a
carbonation/calcination cycle of CaO/CaCO3, Chem. Eng. J. 90 (2002) 303306.
[23] G.S. Grasa, J.C. Abanades, CO2 capture capacity of CaO in long series of
carbonation/calcination cycles, Ind. Eng. Chem. Res. 45 (2006) 88468851.
[24] P. Sun, J. Lim, J.R. Grace, Cyclic CO2 capture by limestone-derived sorbent during
prolonged calcination/carbonation cycling, AIChE J. 54 (2008) 16681677.
[25] L. Zhang, Y. Lu, M. Rostam-Abadi, Sintering of calcium oxide (CaO) during CO2
chemisorption: a reactive molecular dynamics study, Phys. Chem. Chem.
Phys. 14 (2012) 1663316643.
[26] A.I. Lysikov, A.N. Salanov, A.G. Okunev, Change of CO2 carrying capacity of
CaO in isothermal recarbonation-decomposition cycles, Ind. Eng. Chem. Res.
46 (2007) 46334638.
[27] B. Dou, Y. Song, Y. Liu, C. Feng, High temperature CO2 capture using calcium
oxide sorbent in a xed-bed reactor, J. Hazard. Mater. 183 (2010) 759765.
[28] K.R. Rout, J. Fermoso, D. Chen, H.A. Jakobsen, Kinetic rate of CO2 uptake of a
synthetic Ca-based sorbent: experimental data and numerical simulations,
Fuel 120 (2014) 5365.
[29] M. Mohammadi, P. Lahijani, A.R. Mohamed, Refractory dopant-incorporated
CaO from waste eggshell as sustainable sorbent for CO2 capture:
experimental and kinetic studies, Chem. Eng. J. 243 (2014) 455464.
[30] D. Alvarez, J.C. Abanades, Determination of the critical product layer
thickness in the reaction of CaO with CO2, Ind. Eng. Chem. Res. 44 (2005)
56085615.
[31] E. Mostafavi, M.H. Sedghkerdar, N. Mahinpey, Thermodynamic and kinetic
study of CO2 capture with calcium based sorbents: experiments and
modeling, Ind. Eng. Chem. Res. 52 (2013) 47254733.
[32] A. Bandi, M. Specht, P. Sichler, N. Nicoloso, In situ gas conditioning in fuel
reforming for hydrogen generation, in: 5th International Symposium on Gas
Cleaning at High Temperature. US DOE National Energy Technology
Laboratory, Morgantown, USA, 2002.
[33] H. Gupta, L.-S. Fan, Carbonation-calcination cycle using high reactivity
calcium oxide for carbon dioxide separation from ue gas, Ind. Eng. Chem.
Res. 41 (2002) 40354042.
[34] A.J. Nieto-Sanchez, M. Olivares-Marin, S. Garcia, C. Pevida, E.M. CuerdaCorrea, Inuence of the operation conditions on CO2 capture by CaO-derived
sorbents prepared from synthetic CaCO3, Chemosphere 93 (2013) 2148
2158.
[35] M.J. Hsu, K.H. Lee, Y.P. Chyou, CO2 capture at high temperature using calciumbased sorbents, J. Chin. Inst. Eng. 37 (2014) 152164.
[36] V. Materic, M. Hyland, M.I. Jones, B. Northover, High temperature carbonation
of Ca(OH)2: the effect of particle surface area and pore volume, Ind. Eng.
Chem. Res. 53 (2014) 29943000.
[37] S.F. Wu, T.H. Beum, J.I. Yang, J.N. Kim, Properties of Ca-base CO2 sorbent using
Ca(OH)2 as precursor, Ind. Eng. Chem. Res. 46 (2007) 78967899.
[38] G. Grasa, B. GonzAlez, M. Alonso, J.C. Abanades, Comparison of CaO-based
synthetic CO2 sorbents under realistic calcination conditions, Energy Fuels 21
(2007) 35603562.
[39] W.Q. Liu, N.W.L. Low, B. Feng, G.X. Wang, J.C.D. da Costa, Calcium precursors
for the production of CaO sorbents for multicycle CO2 capture, Environ. Sci.
Technol. 44 (2010) 841847.
[40] L. Yang, H. Yu, S. Wang, H. Wang, Q. Zhou, Carbon dioxide captured from ue
gas by modied Ca-based sorbents in xed-bed reactor at high temperature,
Chin. J. Chem. Eng. 21 (2013) 199204.
[41] H. Lu, A. Khan, P.G. Smirniotis, Relationship between structural properties
and CO2 capture performance of CaO-based sorbents obtained from different
organometallic precursors, Ind. Eng. Chem. Res. 47 (2008) 62166220.
[42] H. Lu, P.G. Smirniotis, F.O. Ernst, S.E. Pratsinis, Nanostructured Ca-based
sorbents with high CO2 uptake efciency, Chem. Eng. Sci. 64 (2009) 1936
1943.
[43] A. Silaban, M. Narcida, D.P. Harrison, Calcium acetate as a sorbent precursor
for the removal of carbon dioxide from gas streams at high temperature,
Resour. Conserv. Recycl. 7 (1992) 139153.
[44] N.H. Florin, A.T. Harris, Reactivity of CaO derived from nano-sized CaCO3
particles through multiple CO2 capture-and-release cycles, Chem. Eng. Sci. 64
(2009) 187191.
[45] H. Lu, A. Khan, S.E. Pratsinis, P.G. Smirniotis, Flame-made durable doped-CaO
nanosorbents for CO2 capture, Energy Fuels 23 (2009) 10931100.
[46] C. Luo, Y. Zheng, C. Zheng, J. Yin, C. Qin, B. Feng, Manufacture of calcium-based
sorbents for high temperature cyclic CO2 capture via a solgel process, Int. J.
Greenhouse Gas Control 12 (2013) 193199.
[47] R. Barker, The reactivity of calcium oxide towards carbon dioxide and its use
for energy storage, J. Appl. Chem. Biotechnol. 24 (1974) 221227.
[48] N.H. Florin, A.T. Harris, Screening CaO-based sorbents for CO2 capture in
biomass gasiers, Energy Fuels 22 (2008) 27342742.
[49] E.T. Santos, C. Alfonsn, A.J.S. Chambel, A. Fernandes, A.P. Soares Dias, C.I.C.
Pinheiro, M.F. Ribeiro, Investigation of a stable synthetic solgel CaO sorbent
for CO2 capture, Fuel 94 (2012) 624628.

442

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

[50] J.C. Abanades, D. Alvarez, Conversion limits in the reaction of CO2 with lime,
Energy Fuels 17 (2003) 308315.
[51] Z. Zhou, P. Xu, M. Xie, Z. Cheng, W. Yuan, Modeling of the carbonation kinetics
of a synthetic CaO-based sorbent, Chem. Eng. Sci. 95 (2013) 283290.
[52] F.-Q. Liu, W.-H. Li, B.-C. Liu, R.-X. Li, Synthesis, characterization, and high
temperature CO2 capture of new CaO based hollow sphere sorbents, J. Mater.
Chem. A 1 (2013) 80378044.
[53] J. Park, K.B. Yi, Effects of preparation method on cyclic stability and CO2
absorption capacity of synthetic CaOAMgO absorbent for sorption-enhanced
hydrogen production, Int. J. Hydrogen Energy 37 (2012) 95102.
[54] K.O. Albrecht, K.S. Wagenbach, J.A. Satrio, B.H. Shanks, T.D. Wheelock,
Development of a CaO-based CO2 sorbent with improved cyclic stability,
Ind. Eng. Chem. Res. 47 (2008) 78417848.
[55] W.Q. Liu, J.J. Yin, C.L. Qin, B. Feng, M.H. Xu, Synthesis of CaO-based sorbents
for CO2 capture by a spray-drying technique, Environ. Sci. Technol. 46 (2012)
1126711272.
[56] S.F. Wu, L.L. Wang, Improvement of the stability of a ZrO2-modied Ni-nanoCaO sorption complex catalyst for ReSER hydrogen production, Int. J.
Hydrogen Energy 35 (2010) 65186524.
[57] C.T. Yu, W.C. Chen, Preparation, characterization of Ca/Al carbonate pellets
with TiO2 binder and CO2 sorption at elevated-temperature conditions,
Powder Technol. 239 (2013) 492498.
[58] C.W. Lai, Modication of one-dimensional TiO2 nanotubes with CaO dopants
for high CO2 adsorption, Int. J. Photoenergy 2014 (2014) 9.
[59] J.M. Valverde, A. Perejon, L.A. Perez-Maqueda, Enhancement of fast CO2
capture by a nano-SiO2/CaO composite at Ca-looping conditions, Environ. Sci.
Technol. 46 (2012) 64016408.
[60] Y. Li, C. Zhao, Q. Ren, L. Duan, H. Chen, X. Chen, Effect of rice husk ash addition
on CO2 capture behavior of calcium-based sorbent during calcium looping
cycle, Fuel Process. Technol. 90 (2009) 825834.
[61] M.H. Sedghkerdar, N. Mahinpey, Z.K. Sun, S. Kaliaguine, Novel synthetic sol
gel CaO based pellets using porous mesostructured silica in cyclic CO2 capture
process, Fuel 127 (2014) 101108.
[62] X.Y. Zhang, Z.G. Li, Y. Peng, W.K. Su, X.X. Sun, J.H. Li, Investigation on a novel
CaOAY2O3 sorbent for efcient CO2 mitigation, Chem. Eng. J. 243 (2014) 297
304.
[63] Z.M. Zhou, Y. Qi, M.M. Xie, Z.M. Cheng, W.K. Yuan, Synthesis of CaO-based
sorbents through incorporation of alumina/aluminate and their CO2 capture
performance, Chem. Eng. Sci. 74 (2012) 172180.
[64] M.M. Zhang, Y.X. Peng, Y.Z. Sun, P. Li, J.G. Yu, Preparation of CaOAAl2O3
sorbent and CO2 capture performance at high temperature, Fuel 111 (2013)
636642.
[65] Z.S. Li, N.S. Cai, Y.Y. Huang, Effect of preparation temperature on cyclic CO2
capture and multiple carbonation-calcination cycles for a new Ca-based CO2
sorbent, Ind. Eng. Chem. Res. 45 (2006) 19111917.
[66] C.S. Martavaltzi, A.A. Lemonidou, Development of new CaO based sorbent
materials for CO2 removal at high temperature, Microporous Mesoporous
Mater. 110 (2008) 119127.
[67] S.F. Wu, Q.H. Li, J.N. Kim, K.B. Yi, Properties of a nano CaO/Al2O3 CO2 sorbent,
Ind. Eng. Chem. Res. 47 (2008) 180184.
[68] R. Pacciani, C.R. Muller, J.F. Davidson, J.S. Dennis, A.N. Hayhurst, Synthetic Cabased solid sorbents suitable for capturing CO2 in a uidized bed, Can. J.
Chem. Eng. 86 (2008) 356366.
[69] R. Koirala, G.K. Reddy, P.G. Smirniotis, Single nozzle ame-made highly
durable metal doped Ca-based sorbents for CO2 capture at high temperature,
Energy Fuels 26 (2012) 31033109.
[70] J.S. Dennis, R. Pacciani, The rate and extent of uptake of CO2 by a synthetic,
CaO-containing sorbent, Chem. Eng. Sci. 64 (2009) 21472157.
[71] H.R. Radfarnia, M.C. Iliuta, Metal oxide-stabilized calcium oxide CO2 sorbent
for multicycle operation, Chem. Eng. J. 232 (2013) 280289.
[72] S. Stendardo, L.K. Andersen, C. Herce, Self-activation and effect of
regeneration conditions in CO2-carbonate looping with CaOACa12Al14O33
sorbent, Chem. Eng. J. 220 (2013) 383394.
[73] M. Sayyah, B.R. Ito, M. Rostam-Abadi, Y. Lu, K.S. Suslick, CaO-based sorbents
for CO2 capture prepared by ultrasonic spray pyrolysis, RSC Adv. 3 (2013)
1987219875.
[74] J.N. Kim, C.H. Ko, K.B. Yi, Sorption enhanced hydrogen production using onebody CaOACa12Al14O33ANi composite as catalytic absorbent, Int. J. Hydrogen
Energy 38 (2013) 60726078.
[75] A.M. Kierzkowska, L.V. Poulikakos, M. Broda, C.R. Muller, Synthesis of
calcium-based, Al2O3-stabilized sorbents for CO2 capture using a coprecipitation technique, Int. J. Greenhouse Gas Control 15 (2013) 4854.
[76] C.S. Martavaltzi, A.A. Lemonidou, Parametric study of the CaOACa12Al14O33
synthesis with respect to high CO2 sorption capacity and stability on
multicycle operation, Ind. Eng. Chem. Res. 47 (2008) 95379543.
[77] S.D. Angeli, C.S. Martavaltzi, A.A. Lemonidou, Development of a novelsynthesized Ca-based CO2 sorbent for multicycle operation: Parametric study
of sorption, Fuel 127 (2014) 6269.
[78] H.R. Radfarnia, A. Sayari, A highly efcient CaO-based CO2 sorbent prepared
by a citrate-assisted solgel technique, Chem. Eng. J. 262 (2015) 913920.
[79] M. Broda, C.R. Mller, Synthesis of highly efcient, Ca-based, Al2O3-stabilized,
carbon gel-templated CO2 sorbents, Adv. Mater. 24 (2012) 30593064.
[80] N.H. Florin, J. Blamey, P.S. Fennell, Synthetic CaO-based sorbent for CO2
capture from large-point sources, Energy Fuels 24 (2010) 45984604.
[81] P.-H. Chang, Y.-P. Chang, S.-Y. Chen, C.-T. Yu, Y.-P. Chyou, Ca-rich CaAAloxide, high-temperature-stable sorbents prepared from hydrotalcite

[82]

[83]

[84]

[85]

[86]

[87]

[88]
[89]

[90]

[91]

[92]

[93]
[94]

[95]

[96]

[97]

[98]
[99]

[100]

[101]

[102]
[103]

[104]
[105]
[106]

[107]

[108]

[109]

[110]

precursors: synthesis, characterization, and CO2 capture capacity,


ChemSusChem 4 (2011) 18441851.
P.H. Chang, T.J. Lee, Y.P. Chang, S.Y. Chen, CO2 sorbents with scaffold-like
CaAAl layered double hydroxides as precursors for CO2 capture at high
temperatures, ChemSusChem 6 (2013) 10761083.
P.H. Chang, Y.P. Chang, Y.H. Lai, S.Y. Chen, C.T. Yu, Y.P. Chyou, Synthesis,
characterization and high temperature CO2 capture capacity of nanoscale Cabased layered double hydroxides via reverse microemulsion, J. Alloy Compd.
586 (2014) S498S505.
C.T. Yu, W.C. Chen, Hydrothermal preparation of calciumAaluminum
carbonate sorbent for high-temperature CO2 capture in xed-bed reactor,
Fuel 122 (2014) 179185.
R. Filitz, A.M. Kierzkowska, M. Broda, C.R. Mller, Highly efcient CO2
sorbents: development of synthetic, calcium-rich dolomites, Environ. Sci.
Technol. 46 (2011) 559565.
L.Y. Li, D.L. King, Z.M. Nie, C. Howard, Magnesia-stabilized calcium oxide
absorbents with improved durability for high temperature CO2 capture, Ind.
Eng. Chem. Res. 48 (2009) 1060410613.
W.Q. Liu, B. Feng, Y.Q. Wu, G.X. Wang, J. Barry, J.C.D. da Costa, Synthesis of
sintering-resistant sorbents for CO2 capture, Environ. Sci. Technol. 44 (2010)
30933097.
P.Q. Lan, S.F. Wu, Synthesis of a porous nano-CaO/MgO-based CO2 adsorbent,
Chem. Eng. Technol. 37 (2014) 580586.
R. Koirala, K.R. Gunugunuri, S.E. Pratsinis, P.G. Smirniotis, Effect of zirconia
doping on the structure and stability of CaO-based sorbents for CO2 capture
during extended operating cycles, J. Phys. Chem. C 115 (2011) 2480424812.
H.R. Radfarnia, M.C. Iliuta, Development of zirconium-stabilized calcium
oxide absorbent for cyclic high-temperature CO2 capture, Ind. Eng. Chem. Res.
51 (2012) 1039010398.
G.K. Reddy, S. Quillin, P. Smirniotis, Inuence of the synthesis method on the
structure and CO2 adsorption properties of Ca/Zr sorbents, Energy Fuels 28
(2014) 32923299.
M. Zhao, M. Bilton, A.P. Brown, A.M. Cunliffe, E. Dvininov, V. Dupont, T.P.
Comyn, S.J. Milne, Durability of CaOACaZrO3 sorbents for high-temperature
CO2 capture prepared by a wet chemical method, Energy Fuels 28 (2014)
12751283.
M. Broda, C.R. Muller, Solgel-derived, CaO-based, ZrO2-stabilized CO2
sorbents, Fuel 127 (2014) 94100.
Z.S. Li, Y. Liu, N.S. Cai, Understanding the enhancement effect of hightemperature steam on the carbonation reaction of CaO with CO2, Fuel 127
(2014) 8893.
M. Broda, V. Manovic, E.J. Anthony, C.R. Muller, Effect of pelletization and
addition of steam on the cyclic performance of carbon-templated, CaO-based
CO2 sorbents, Environ. Sci. Technol. 48 (2014) 53225328.
S. Champagne, D.Y. Lu, A. Macchi, R.T. Symonds, E.J. Anthony, Inuence of
steam injection during calcination on the reactivity of CaO-based sorbent for
carbon capture, Ind. Eng. Chem. Res. 52 (2013) 22412246.
N. Rong, Q.H. Wang, M.X. Fang, L.M. Cheng, Z.Y. Luo, K.F. Cen, Steam hydration
reactivation of CaO-based sorbent in cyclic carbonation/calcination for CO2
capture, Energy Fuels 27 (2013) 53325340.
V. Manovic, E.J. Anthony, Carbonation of CaO-based sorbents enhanced by
steam addition, Ind. Eng. Chem. Res. 49 (2010) 91059110.
R.T. Symonds, D.Y. Lu, R.W. Hughes, E.J. Anthony, A. Macchi, CO2 capture from
simulated syngas via cyclic carbonation/calcination for a naturally occurring
limestone: pilot-plant testing, Ind. Eng. Chem. Res. 48 (2009) 84318440.
Y. Wang, S. Lin, Y. Suzuki, Limestone calcination with CO2 capture (II):
decomposition in CO2/steam and CO2/N2 atmospheres, Energy Fuels 22
(2008) 23262331.
R.T. Symonds, D.Y. Lu, A. Macchi, R.W. Hughes, E.J. Anthony, CO2 capture from
syngas via cyclic carbonation/calcination for a naturally occurring limestone:
modelling and bench-scale testing, Chem. Eng. Sci. 64 (2009) 35363543.
S. Yang, Y. Xiao, Steam catalysis in CaO carbonation under low steam partial
pressure, Ind. Eng. Chem. Res. 47 (2008) 40434048.
R.Y. Sun, Y.J. Li, C.T. Liu, X. Xie, C.M. Lu, Utilization of lime mud from paper
mill as CO2 sorbent in calcium looping process, Chem. Eng. J. 221 (2013) 124
132.
V. Manovic, E.J. Anthony, Steam reactivation of spent CaO-based sorbent for
multiple CO2 capture cycles, Environ. Sci. Technol. 41 (2007) 14201425.
F. Zeman, Effect of steam hydration on performance of lime sorbent for CO2
capture, Int. J. Greenhouse Gas Control 2 (2008) 203209.
A. Coppola, P. Salatino, F. Montagnaro, F. Scala, Reactivation by water
hydration of the CO2 capture capacity of a calcium looping sorbent, Fuel 127
(2014) 109115.
V. Materic, M. Hyland, M.I. Jones, R. Holt, Investigation of the friability of Ca
looping sorbents during and after hydration based reactivation, Fuel 127
(2014) 7077.
I. Martinez, G. Grasa, R. Murillo, B. Arias, J.C. Abanades, Evaluation of CO2
carrying capacity of reactivated CaO by hydration, Energy Fuels 25 (2011)
12941301.
C.-C. Li, J.-Y. Cheng, W.-H. Liu, C.-M. Huang, H.-W. Hsu, H.-P. Lin,
Enhancement in cyclic stability of the CO2 adsorption capacity of CaObased sorbents by hydration for the calcium looping cycle, J. Taiwan Inst.
Chem. Eng. 45 (2014) 227232.
V. Manovic, E.J. Anthony, Thermal activation of CaO-based sorbent and selfreactivation during CO2 capture looping cycles, Environ. Sci. Technol. 42
(2008) 41704174.

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444


[111] Z. Chen, H.S. Song, M. Portillo, C.J. Lim, J.R. Grace, E.J. Anthony, Long-term
calcination/carbonation cycling and thermal pretreatment for CO2 capture by
limestone and dolomite, Energy Fuels 23 (2009) 14371444.
[112] B. Arias, J.C. Abanades, E.J. Anthony, Model for self-reactivation of highly
sintered CaO particles during CO2 capture looping cycles, Energy Fuels 25
(2011) 19261930.
[113] J.M. Valverde, P.E. Sanchez-Jimenez, A. Perejon, L.A. Perez-Maqueda, Role of
looping-calcination conditions on self-reactivation of thermally pretreated
CO2 sorbents based on CaO, Energy Fuels 27 (2013) 33733384.
[114] P.E. Sanchez-Jimenez, J.M. Valverde, L.A. Perez-Maqueda, Multicyclic
conversion of limestone at Ca-looping conditions: the role of solid-state
diffusion controlled carbonation, Fuel 127 (2014) 131140.
[115] G. Grasa, I. Martinez, M.E. Diego, J.C. Abanades, Determination of CaO
carbonation kinetics under recarbonation conditions, Energy Fuels 28 (2014)
40334042.
[116] J.M. Valverde, P.E. Sanchez-Jimenez, L.A. Perez-Maqueda, High and stable CO2
capture capacity of natural limestone at Ca-looping conditions by heat
pretreatment and recarbonation synergy, Fuel 123 (2014) 7985.
[117] J.M. Valverde, P.E. Sanchez-Jimenez, L.A. Perez-Maqueda, Effect of heat
pretreatment/recarbonation in the Ca-looping process at realistic calcination
conditions, Energy Fuels 28 (2014) 40624067.
[118] Y. Li, C. Zhao, C. Qu, L. Duan, Q. Li, C. Liang, CO2 capture using CaO modied
with ethanol/water solution during cyclic calcination/carbonation, Chem.
Eng. Technol. 31 (2008) 237244.
[119] S.P. Wang, H. Shen, S.S. Fan, Y.J. Zhao, X.B. Ma, J.L. Gong, Enhanced CO2
adsorption capacity and stability using CaO-based adsorbents treated by
hydration, AIChE J. 59 (2013) 35863593.
[120] R. Sun, Y. Li, S. Wu, C. Liu, H. Liu, C. Lu, Enhancement of CO2 capture capacity by
modifying limestone with propionic acid, Powder Technol. 233 (2013) 814.
[121] Y. Li, C. Zhao, H. Chen, C. Liang, L. Duan, W. Zhou, Modied CaO-based sorbent
looping cycle for CO2 mitigation, Fuel 88 (2009) 697704.
[122] Y. Li, R. Sun, H. Liu, C. Lu, Cyclic CO2 capture behavior of limestone modied
with pyroligneous acid (PA) during calcium looping cycles, Ind. Eng. Chem.
Res. 50 (2011) 1022210228.
[123] H.R. Radfarnia, M.C. Iliuta, Limestone acidication using citric acid coupled
with two-step calcination for improving the CO2 sorbent activity, Ind. Eng.
Chem. Res. 52 (2013) 70027013.
[124] F.N. Ridha, V. Manovic, A. Macchi, M.A. Anthony, E.J. Anthony, Assessment of
limestone treatment with organic acids for CO2 capture in Ca-looping cycles,
Fuel Process. Technol. 116 (2013) 284291.
[125] V. Manovic, E.J. Anthony, Screening of binders for pelletization of CaO-based
sorbents for CO2 capture, Energy Fuels 23 (2009) 47974804.
[126] F.N. Ridha, V. Manovic, Y.H. Wu, A. Macchi, E.J. Anthony, Pelletized CaO-based
sorbents treated with organic acids for enhanced CO2 capture in Ca-looping
cycles, Int. J. Greenhouse Gas Control 17 (2013) 357365.
[127] H. Chen, C. Zhao, Y. Yang, Enhancement of attrition resistance and cyclic CO2
capture of calcium-based sorbent pellets, Fuel Process. Technol. 116 (2013)
116122.
[128] E. Ochoa-Fernandez, C. Lacalle-Vila, T. Zhao, M. Ronning, D. Chen,
Experimental demonstration of H2 production by CO2 sorption enhanced
steam methane reforming using ceramic acceptors, Stud. Surf. Sci. Catal. 167
(2007) 159164.
[129] J. Ida, R.T. Xiong, Y.S. Lin, Synthesis and CO2 sorption properties of pure and
modied lithium zirconate, Sep. Purif. Technol. 36 (2004) 4151.
[130] G. Pannocchia, M. Puccini, M. Seggiani, S. Vitolo, Experimental and modeling
studies on high-temperature capture of CO2 using lithium zirconate based
sorbents, Ind. Eng. Chem. Res. 46 (2007) 66966706.
[131] M. Kato, K. Nakagawa, K. Essaki, Y. Maezawa, S. Takeda, R. Kogo, Y. Hagiwara,
Novel CO2 absorbents using lithium-containing oxide, Int. J. Appl. Ceram.
Technol. 2 (2005) 467475.
[132] D.J. Fauth, E.A. Frommell, J.S. Hoffman, R.P. Reasbeck, H.W. Pennline, Eutectic
salt promoted lithium zirconate: novel high temperature sorbent for CO2
capture, Fuel Process. Technol. 86 (2005) 15031521.
[133] H. Pfeiffer, P. Bosch, Thermal stability and high-temperature carbon dioxide
sorption on hexa-lithium zirconate (Li6Zr2O7), Chem. Mater. 17 (2005) 1704
1710.
[134] K.B. Yi, D.O. Eriksen, Low temperature liquid state synthesis of lithium
zirconate and its characteristics as a CO2 sorbent, Sep. Sci. Technol. 41 (2006)
283296.
[135] T. Ohashi, K. Nakagawa, Effect of potassium carbonate additive on CO2
absorption in lithium zirconate powder, in: S.M. Kauzlarich, E.M. McCarron,
A.W. Sleight, H.C. zurLoye (Eds.) Solid-State Chemistry of Inorganic Materials
II, 1999, pp. 249254.
[136] J. Ida, Y.S. Lin, Mechanism of high-temperature CO2 sorption on lithium
zirconate, Environ. Sci. Technol. 37 (2003) 19992004.
[137] E. Ochoa-Fernandez, G. Haugen, T. Zhao, M. Ronning, I. Aartun, B. Borresen, E.
Rytter, M. Ronnekleiv, D. Chen, Process design simulation of H2 production by
sorption enhanced steam methane reforming: evaluation of potential CO2
acceptors, Green Chem. 9 (2007) 654662.
[138] E. Ochoa-Fernandez, M. Ronning, X. Yu, T. Grande, D. Chen, Compositional
effects of nanocrystalline lithium zirconate on its CO2 capture properties, Ind.
Eng. Chem. Res. 47 (2008) 434442.
[139] C. Wang, B. Dou, Y. Song, H. Chen, Y. Xu, B. Xie, High temperature CO2
sorption on Li2ZrO3 based sorbents, Ind. Eng. Chem. Res. 53 (2014) 12744
12752.

443

[140] R.T. Xiong, J. Ida, Y.S. Lin, Kinetics of carbon dioxide sorption on potassiumdoped lithium zirconate, Chem. Eng. Sci. 58 (2003) 43774385.
[141] B.N. Nair, T. Yamaguchi, H. Kawamura, S.I. Nakao, K. Nakagawa, Processing of
lithium zirconate for applications in carbon dioxide separation: structure and
properties of the powders, J. Am. Ceram. Soc. 87 (2004) 6874.
[142] A. Iwan, H. Stephenson, W.C. Ketchie, A.A. Lapkin, High temperature
sequestration of CO2 using lithium zirconates, Chem. Eng. J. 146 (2009)
249258.
[143] Q. Xiao, Y. Liu, Y. Zhong, W. Zhu, A citrate solgel method to synthesize
Li2ZrO3 nanocrystals with improved CO2 capture properties, J. Mater. Chem.
21 (2011) 38383842.
[144] Q. Xiao, X. Tang, Y. Liu, Y. Zhong, W. Zhu, Citrate route to prepare K-doped
Li2ZrO3 sorbents with excellent CO2 capture properties, Chem. Eng. J. 174
(2011) 231235.
[145] H.R. Radfarnia, M.C. Iliuta, Surfactant-template/ultrasound-assisted method
for the preparation of porous nanoparticle lithium zirconate, Ind. Eng. Chem.
Res. 50 (2011) 92959305.
[146] M.J. Venegas, E. Fregoso-Israel, R. Escamilla, H. Pfeiffer, Kinetic and reaction
mechanism of CO2 sorption on Li4SiO4: study of the particle size effect, Ind.
Eng. Chem. Res. 46 (2007) 24072412.
[147] I.C. Romero-Ibarra, J. Ortiz-Landeros, H. Pfeiffer, Microstructural and CO2
chemisorption analyses of Li4SiO4: effect of surface modication by the ball
milling process, Thermochim. Acta 567 (2013) 118124.
[148] S.Y. Shan, Q.M. Jia, L.H. Jiang, Q.C. Li, Y.M. Wang, J.H. Peng, Novel Li4SiO4based sorbents from diatomite for high temperature CO2 capture, Ceram. Int.
39 (2013) 54375441.
[149] K. Wang, P.F. Zhao, X. Guo, Y.M. Li, D.T. Han, Y. Chao, Enhancement of
reactivity in Li4SiO4-based sorbents from the nano-sized rice husk ash for
high-temperature CO2 capture, Energy Convers. Manage. 81 (2014) 447454.
[150] M. Seggiani, M. Puccini, S. Vitolo, Alkali promoted lithium orthosilicate for
CO2 capture at high temperature and low concentration, Int. J. Greenhouse
Gas Control 17 (2013) 2531.
[151] H. Pfeiffer, E. Lima, P. Bosch, Lithium-sodium metazirconate solid solutions,
Li2-xNaxZrO3 (0 <= x <= 2): a hierarchical architecture, Chem. Mater. 18 (2006)
26422647.
[152] H. Pfeiffer, C. Vazquez, V.H. Lara, P. Bosch, Thermal behavior and CO2
absorption of Li2-xNaxZrO3 solid solutions, Chem. Mater. 19 (2007) 922926.
[153] A. Sandoval-Diaz, H. Pfeiffer, Effects of potassium doping on the composition,
structure and carbon dioxide chemisorption of Na2ZrO3, Rev. Mex. Fis. 54
(2008) 6568.
[154] V. Guzman-Velderrain, D. Delgado-Vigil, V. Collins-Martinez, A. Lopez Ortiz,
Synthesis, characterization and evaluation of sodium doped lithium zirconate
as a high temperature CO2 absorbent, J. New Mater. Electrochem. Syst. 11
(2008) 131136.
[155] L.O. Gamboa Hernandez, D. Lardizabal Gutierrez, V. Collins-Martinez, A.
Lopez Ortiz, Synthesis characterization and high temperature CO2 capture
evaluation of Li2ZrO3ANa2ZrO3 mixtures, J. New Mater. Electrochem. Syst. 11
(2008) 137142.
[156] H.R. Radfarnia, M.C. Iliuta, Application of surfactant-template technique for
preparation of sodium zirconate as high temperature CO2 sorbent, Sep. Purif.
Technol. 93 (2012) 98106.
[157] L. Martinez-dlCruz, H. Pfeiffer, Microstructural thermal evolution of the
Na2CO3 phase produced during a Na2ZrO3ACO2 chemisorption process, J.
Phys. Chem. C 116 (2012) 96759680.
[158] T. valos-Rendn, J. Casa-Madrid, H. Pfeiffer, Thermochemical capture of
carbon dioxide on lithium aluminates (LiAlO2 and Li5AlO4): a new option for
the CO2 absorption, J. Phys. Chem. A 113 (2009) 69196923.
[159] T. valos-Rendn, V.H. Lara, H. Pfeiffer, CO2 chemisorption and cyclability
analyses of lithium aluminate polymorphs (a-and b-Li5AlO4), Ind. Eng. Chem.
Res. 51 (2012) 26222630.
[160] L.M. Palacios-Romero, H. Pfeiffer, Lithium cuprate (Li2CuO2): a new possible
ceramic material for CO2 chemisorption, Chem. Lett. 37 (2008) 862863.
[161] Y. Matsukura, T. Okumura, R. Kobayashi, K. Oh-ishi, Synthesis and CO2
absorption properties of single-phase Li2CuO2 as a CO2 absorbent, Chem. Lett.
39 (2010) 966967.
[162] I. Yanase, A. Kameyama, H. Kobayashi, CO2 absorption and structural phase
transition of a-LiFeO2, J. Ceram. Soc. Jpn. 118 (2010) 4851.
[163] M. Kato, K. Essaki, K. Nakagawa, Y. Suyama, K. Terasaka, CO2 absorption
properties of lithium ferrite for application as a high-temperature CO2
absorbent, J. Ceram. Soc. Jpn. 113 (2005) 684686.
[164] F. Durn-Muoz, I.C. Romero-Ibarra, H. Pfeiffer, Analysis of the CO2
chemisorption reaction mechanism in lithium oxosilicate (Li8SiO6): a new
option for high-temperature CO2 capture, J. Mater. Chem. A 1 (2013) 3919
3925.
[165] Y. Duan, H. Pfeiffer, B. Li, I.C. Romero-Ibarra, D.C. Sorescu, D.R. Luebke, J.W.
Halley, CO2 capture properties of lithium silicates with different ratios of
Li2O/SiO2: an ab initio thermodynamic and experimental approach, Phys.
Chem. Chem. Phys. 15 (2013) 1353813558.
[166] I.C. Romero-Ibarra, F. Durn-Muoz, H. Pfeiffer, Inuence of the K-, Na- and
K-Na-carbonate additions during the CO2 chemisorption on lithium
oxosilicate (Li8SiO6), Greenhouse Gases Sci. Technol. 4 (2014) 145154.
[167] N. Togashi, T. Okumura, K. Oh-ishi, Synthesis and CO2 absorption property of
Li4TiO4 as a novel CO2 absorbent, J. Ceram. Soc. Jpn. 115 (2007) 324328.
[168] S. Ueda, R. Inoue, K. Sasaki, K. Wakuta, T. Ariyama, CO2 absorption and
desorption abilities of Li2OATiO2 compounds, ISIJ Int. 51 (2011) 530537.

444

M. Shokrollahi Yancheshmeh et al. / Chemical Engineering Journal 283 (2016) 420444

[169] P. Sanchez-Camacho, I.C. Romero-Ibarra, Y. Duan, H. Pfeiffer, Thermodynamic


and kinetic analyses of the CO2 chemisorption mechanism on Na2TiO3:
experimental and theoretical evidences, J. Phys. Chem. C 118 (2014) 19822
19832.
[170] F. Fujishiro, K. Fukasawa, T. Hashimoto, CO2 absorption and desorption
properties of single phase Ba2Fe2O5 and analysis of their mechanism using
thermodynamic calculation, J. Am. Ceram. Soc. 94 (2011) 36753678.
[171] F. Fujishiro, Y. Kojima, T. Hashimoto, Kinetics and mechanism of chemical
reaction of CO2 and Ba2Fe2O5 under various CO2 partial pressures, J. Am.
Ceram. Soc. 95 (2012) 36343637.
[172] K. Essaki, M. Kato, K. Nakagawa, CO2 removal at high temperature using
packed bed of lithium silicate pellets, J. Ceram. Soc. Jpn. 114 (2006) 739742.
[173] V.L. Meja-Trejo, E. Fregoso-Israel, H. Pfeiffer, Textural, structural, and CO2
chemisorption effects produced on the lithium orthosilicate by its doping
with sodium (Li4-xNaxSiO4), Chem. Mater. 20 (2008) 71717176.
[174] Z. Qi, H. Daying, L. Yang, Y. Qian, Z. Zibin, Analysis of CO2 sorption/desorption
kinetic behaviors and reaction mechanisms on Li4SiO4, AIChE J. 59 (2013)
901911.
[175] I. Alcrreca-Corte, E. Fregoso-Israel, H. Pfeiffer, CO2 absorption on Na2ZrO3: a
kinetic analysis of the chemisorption and diffusion processes, J. Phys. Chem. C
112 (2008) 65206525.
[176] J.D. Holladay, J. Hu, D.L. King, Y. Wang, An overview of hydrogen production
technologies, Catal. Today 139 (2009) 244260.
[177] R. Kothari, D. Buddhi, R.L. Sawhney, Comparison of environmental and
economic aspects of various hydrogen production methods, Renewable
Sustainable Energy Rev. 12 (2008) 553563.
[178] S.S. Penner, Steps toward the hydrogen economy, Energy 31 (2006) 3343.
[179] R.K. Dixon, Advancing towards a hydrogen energy economy: status,
opportunities and barriers, Mitigation Adaptation Strategies Global Change
12 (2007) 325341.
[180] J.N. Armor, Catalysis and the hydrogen economy, Catal. Lett. 101 (2005) 131135.
[181] U. Izquierdo, V.L. Barrio, J.F. Cambra, J. Requies, M.B. Guemez, P.L. Arias, G.
Kolb, R. Zapf, A.M. Gutierrez, J.R. Arraibi, Hydrogen production from methane
and natural gas steam reforming in conventional and microreactor reaction
systems, Int. J. Hydrogen Energy 37 (2012) 70267033.
[182] R. Chaubey, S. Sahu, O.O. James, S. Maity, A review on development of
industrial processes and emerging techniques for production of hydrogen
from renewable and sustainable sources, Renewable Sustainable Energy Rev.
23 (2013) 443462.
[183] J.R. Hufton, S. Mayorga, S. Sircar, Sorption-enhanced reaction process for
hydrogen production, AIChD J. 45 (1999) 248256.
[184] K.S. Go, S.R. Son, S.D. Kim, K.S. Kang, C.S. Park, Hydrogen production from
two-step steam methane reforming in a uidized bed reactor, Int. J. Hydrogen
Energy 34 (2009) 13011309.
[185] B.T. Carvill, J.R. Hufton, M. Anand, S. Sircar, Sorption-enhanced reaction
process, AIChE J. 42 (1996) 27652772.
[186] S. Sircar, M. Anand, B.T. Carvill, J.R. Hufton, S. Mayorga, R.N. Miller, Sorption
enhanced reaction process for production of hydrogen, in: Proc. U.S. DOE
Hydrogen Program Review, vol. 1, 1995, p. 815.
[187] K. Johnsen, H.J. Ryu, J.R. Grace, C.J. Lim, Sorption-enhanced steam reforming
of methane in a uidized bed reactor with dolomite as CO2-acceptor, Chem.
Eng. Sci. 61 (2006) 11951202.
[188] Y. Ding, E. Alpay, Adsorption-enhanced steam-methane reforming, Chem.
Eng. Sci. 55 (2000) 39293940.
[189] K.B. Yi, D.P. Harrison, Low-pressure sorption-enhanced hydrogen production,
Ind. Eng. Chem. Res. 44 (2005) 16651669.
[190] A.L. Ortiz, D.P. Harrison, Hydrogen production using sorption-enhanced
reaction, Ind. Eng. Chem. Res. 40 (2001) 51025109.
[191] C.S. Martavaltzi, E.P. Pampaka, E.S. Korkakaki, A.A. Lemonidou, Hydrogen
production via steam reforming of methane with simultaneous CO2 Capture
over CaOACa12Al14O33, Energy Fuels 24 (2010) 25892595.
[192] B. Balasubramanian, A.L. Ortiz, S. Kaytakoglu, D.P. Harrison, Hydrogen from
methane in a single-step process, Chem. Eng. Sci. 54 (1999) 35433552.
[193] Z.S. Li, N.S. Cai, J.B. Yang, Continuous production of hydrogen from sorptionenhanced steam methane reforming in two parallel xed-bed reactors
operated in a cyclic manner, Ind. Eng. Chem. Res. 45 (2006) 87888793.
[194] M. Broda, A.M. Kierzkowska, C.R. Muller, Sorbent-enhanced steam methane
reforming reaction studied over a Ca-based CO2 sorbent and Ni catalyst,
Chem. Eng. Technol. 36 (2013) 14961502.
[195] P. Xu, M.M. Xie, Z.M. Cheng, Z.M. Zhou, CO2 capture performance of CaObased sorbents prepared by a solgel method, Ind. Eng. Chem. Res. 52 (2013)
1216112169.
[196] M. Broda, V. Manovic, Q. Imtiaz, A.M. Kierzkowska, E.J. Anthony, C.R. Muller,
High-purity hydrogen via the sorption-enhanced steam methane reforming
reaction over a synthetic CaO-based sorbent and a Ni catalyst, Environ. Sci.
Technol. 47 (2013) 60076014.
[197] A. Kapil, S.A. Bhat, J. Sadhukhan, Multiscale characterization framework for
sorption enhanced reaction processes, AIChE J. 54 (2008) 10251036.
[198] K.R. Rout, H.A. Jakobsen, A numerical study of pellets having both catalytic
and capture properties for SESMR process: kinetic and product layer diffusion
controlled regimes, Fuel Process. Technol. 106 (2013) 231246.
[199] N. Chanburanasiri, A.M. Ribeiro, A.E. Rodrigues, A. Arpornwichanop, N.
Laosiripojana, P. Praserthdam, S. Assabumrungrat, Hydrogen production via

[200]

[201]

[202]

[203]

[204]

[205]

[206]

[207]

[208]

[209]
[210]

[211]

[212]

[213]
[214]

[215]
[216]

[217]

[218]

[219]

[220]

[221]

[222]

[223]

[224]

[225]

[226]

sorption enhanced steam methane reforming process using Ni/CaO


multifunctional catalyst, Ind. Eng. Chem. Res. 50 (2011) 1366213671.
M.M. Xie, Z.M. Zhou, Y. Qi, Z.M. Cheng, W.K. Yuan, Sorption-enhanced steam
methane reforming by in-situ CO2 capture on a CaOACa9Al6O18 sorbent,
Chem. Eng. J. 207 (2012) 142150.
C.S. Martavaltzi, A.A. Lemonidou, Hydrogen production via sorption
enhanced reforming of methane: Development of a novel hybrid materialreforming catalyst and CO2 sorbent, Chem. Eng. Sci. 65 (2010) 41344140.
H.Z. Feng, P.Q. Lan, S.F. Wu, A study on the stability of a NiOACaO/Al2O3
complex catalyst by La2O3 modication for hydrogen production, Int. J.
Hydrogen Energy 37 (2012) 1416114166.
M. Broda, A.M. Kierzkowska, D. Baudouin, Q. Imtiaz, C. Coperet, C.R. Muller,
Sorbent-enhanced methane reforming over a NiACa-based, bifunctional
catalyst sorbent, ACS Catal. 2 (2012) 16351646.
H.R. Radfarnia, M.C. Iliuta, Development of Al-stabilized CaOnickel hybrid
sorbent-catalyst for sorption-enhanced steam methane reforming, Chem.
Eng. Sci. 109 (2014) 212219.
H.R. Radfarnia, M.C. Iliuta, Hydrogen production by sorption-enhanced steam
methane reforming process using CaOAZr/Ni bifunctional sorbentcatalyst,
Chem. Eng. Process. Process Intensif. 86 (2014) 96103.
L. Barelli, G. Bidini, A. Di Michele, F. Gallorini, C. Petrillo, F. Sacchetti,
Synthesis and test of sorbents based on calcium aluminates for SE-SR, Appl.
Energy 127 (2014) 8192.
K.B. Yi, J. Meyer, D. Eriksen, Selection, preparation and study of high
temperature novel CO2 sorbents for sorption enhanced SMR, in: Nordic H2Seminar, Oslo, 2006.
E. Ochoa-Fernandez, T.J. Zhao, M. Ronning, D. Chen, Effects of steam addition
on the properties of high temperature ceramic CO2 acceptors, J. Environ. Eng.
(ASCE) 135 (2009) 397403.
B. Suresh, S. Schlag, T. Kumamoto, Y. Ring, Hydrogen Report CEH,
<http://chemical.ihs.com/CEH/Public/Reports/743.5000>, 2010.
X. Wang, M. Li, S. Li, H. Wang, S. Wang, X. Ma, Hydrogen production by
glycerol steam reforming with/without calcium oxide sorbent: a comparative
study of thermodynamic and experimental work, Fuel Process. Technol. 91
(2010) 18121818.
L. He, J.M.S. Parra, E.A. Blekkan, D. Chen, Towards efcient hydrogen
production from glycerol by sorption enhanced steam reforming, Energy
Environ. Sci. 3 (2010) 10461056.
J. Fermoso, L. He, D. Chen, Sorption enhanced steam reforming (SESR): a
direct route towards efcient hydrogen production from biomass-derived
compounds, J. Chem. Technol. Biotechnol. 87 (2012) 13671374.
P.D. Vaidya, A.E. Rodrigues, Glycerol reforming for hydrogen production: a
review, Chem. Eng. Technol. 32 (2009) 14631469.
B. Dou, Y. Song, C. Wang, H. Chen, Y. Xu, Hydrogen production from catalytic
steam reforming of biodiesel byproduct glycerol: issues and challenges,
Renewable Sustainable Energy Rev. 30 (2014) 950960.
Y.-C. Lin, Catalytic valorization of glycerol to hydrogen and syngas, Int. J.
Hydrogen Energy 38 (2013) 26782700.
A.C.D. Freitas, R. Guirardello, Comparison of several glycerol reforming
methods for hydrogen and syngas production using Gibbs energy
minimization, Int. J. Hydrogen Energy 39 (2014) 1796917984.
M. Benito, R. Padilla, J.L. Sanz, L. Daza, Thermodynamic analysis and
performance of a 1 kW bioethanol processor for a PEMFC operation, J.
Power Sources 169 (2007) 123130.
F. Daz, F. Alvarado Gracia, Oxidative steam reforming of glycerol for
hydrogen production: Thermodynamic analysis including different carbon
deposits representation and CO2 adsorption, Int. J. Hydrogen Energy 37
(2012) 1482014830.
C. He, J. Zheng, K. Wang, H. Lin, J.-Y. Wang, Y. Yang, Sorption enhanced
aqueous phase reforming of glycerol for hydrogen production over PtANi
supported on multi-walled carbon nanotubes, Appl. Catal. B: Environ. 162
(2015) 401411.
I. Iliuta, H.R. Radfarnia, M.C. Iliuta, Hydrogen production by sorptionenhanced steam glycerol reforming: sorption kinetics and reactor
simulation, AIChE J. 59 (2013) 21052118.
H. Chen, T. Zhang, B. Dou, V. Dupont, P. Williams, M. Ghadiri, Y. Ding,
Thermodynamic analyses of adsorption-enhanced steam reforming of glycerol
for hydrogen production, Int. J. Hydrogen Energy 34 (2009) 72087222.
B. Dou, V. Dupont, G. Rickett, N. Blakeman, P.T. Williams, H. Chen, Y. Ding, M.
Ghadiri, Hydrogen production by sorption-enhanced steam reforming of
glycerol, Bioresour. Technol. 100 (2009) 35403547.
J. Fermoso, L. He, D. Chen, Production of high purity hydrogen by sorption
enhanced steam reforming of crude glycerol, Int. J. Hydrogen Energy 37
(2012) 1404714054.
B. Dou, G.L. Rickett, V. Dupont, P.T. Williams, H. Chen, Y. Ding, M. Ghadiri,
Steam reforming of crude glycerol with in situ CO2 sorption, Bioresour.
Technol. 101 (2010) 24362442.
B. Dou, C. Wang, H. Chen, Y. Song, B. Xie, Continuous sorption-enhanced
steam reforming of glycerol to high-purity hydrogen production, Int. J.
Hydrogen Energy 38 (2013) 1190211909.
B. Dou, Y. Song, C. Wang, H. Chen, M. Yang, Y. Xu, Hydrogen production by
enhanced-sorption chemical looping steam reforming of glycerol in movingbed reactors, Appl. Energy 130 (2014) 342349.

You might also like