Sie sind auf Seite 1von 11

Fuel 136 (2014) 46–56

Contents lists available at ScienceDirect

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

Aspen Plus simulations of fluidised beds for chemical looping


combustion
Rosario Porrazzo, Graeme White, Raffaella Ocone ⇑
Chemical Engineering, Heriot-Watt University, Edinburgh EH14 4AS, UK

h i g h l i g h t s

 Aspen Plus is applied to chemical looping combustion.


 The bubbling bed and the circulating bed are modelled using appropriate combinations of ideal reactors.
 The effect of the shrinking-core model for the non-catalytic heterogeneous reactions is investigated.
 The methodology is checked against experimental data from the literature.

a r t i c l e i n f o a b s t r a c t

Article history: Chemical Looping Combustion (CLC) is a technology able to generate energy whilst managing CO2
Received 1 October 2013 emissions. A system composed by two interconnected fluidised beds is often used in CLC: the two fluid-
Received in revised form 19 June 2014 ised beds are employed for carrying out the oxidation and reduction reactions of the metal oxide
Accepted 24 June 2014
employed as oxygen carrier. In this work, a model to implement fluidised bed systems in Aspen Plus is
Available online 16 July 2014
presented. Depending on the hydrodynamic regimes, two different models are considered: one of the
two fluidised beds, called the fuel reactor, is modelled according to the two-phase theory (i.e. emulsion
Keywords:
and bubble phase) whilst the other bed, called the air reactor, is assumed to operate in the fast fluidisa-
Chemical looping combustion
Aspen Plus
tion regime. Kinetic equations for heterogeneous gas/solid reactions are also considered in the model.
Fluidised beds Simulation tests for each fluidised bed are carried out, and comparisons are made with experimental data
Carbon capture from the literature. A comparison with the largely used Gibbs reactor model is carried out showing the
advantages of using the models developed here. In addition, the net heat duty of the whole process is
calculated and the role of the main variables that affect the process is investigated.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction technologies with economical capture, transportation, and safe


storage solutions.
Carbon capture from power plants is one of the remits of A number of carbon capture techniques can be implemented in
governments worldwide as part of their duty to mitigate climate the operation of fossil fuel based power plants. They can be divided
change. The increase in anthropogenic greenhouse gases is into post-combustion capture, pre-combustion capture, and
believed to be related to the planet temperature rise. CO2 accounts oxy-combustion. Post-combustion capture removes CO2 from the
for up to 64% of emissions of enhanced greenhouse effect [1] con- combustion flue gases whereas in the pre-combustion capture fuel
stituting about 15% of the flue gas stream from coal combustion reacts with oxygen obtained from the air separation unit (ASU), in
power plants, where the total amount of CO2 produced from coal the presence of steam, to produce syngas. The CO in the syngas is
based power plants accounts for more than 40% of all anthropo- converted to CO2 through the water–gas shift reaction (WGS). The
genic CO2 emissions [2]. The target for the future is an energy products of the WGS reaction contain high CO2 and H2 amounts:
mix which involves the development of technologies based on CO2 is captured whilst H2 is used to generate electricity through
nuclear and/or renewable energies with lower emissions of CO2. a combined cycle power generation system. In the oxy-combustion
Meanwhile, a viable energy solution for the reduction of CO2 capture, combustion takes place using oxygen instead of air, to
emissions could consist in coupling fossil fuel energy conversion generate a concentrated CO2 and H2O vapour streams.
Fossil fuel based power plants, using traditional post-combustion
⇑ Corresponding author. separation units, involve large energy losses and additional
E-mail address: r.ocone@hw.ac.uk (R. Ocone). costs for the electricity generated. The existing techniques for

http://dx.doi.org/10.1016/j.fuel.2014.06.053
0016-2361/Ó 2014 Elsevier Ltd. All rights reserved.
R. Porrazzo et al. / Fuel 136 (2014) 46–56 47

Nomenclature

a decay index T operative temperature (°C)


Ab cross sectional area of the bubble phase (m2) Ub bubble superficial gas velocity (m/s)
Ae cross sectional area of the emulsion phase (m2) Ubr bubble superficial gas velocity at minimum fluidisation
At cross sectional area of the whole bed (m2) (m/s)
CA molar concentration of A gas reactant (kmol/m3) Ue emulsion superficial gas velocity (m/s)
DBB diameter of the bubbling bed (m) Umf superficial gas velocity at minimum fluidisation (m/s)
db bubble diameter (m) Uo inlet superficial gas velocity (m/s)
dp particle diameter (m) Ut particle terminal velocity (m/s)
DR diameter of the riser (m) Vb bubble volume (m3)
F component molar flowrate (kmol/s) VCSTR CSTR volume (m3)
Fs inlet solid mass flowrate (kg/s) Ve emulsion volume (m3)
g gravity (m/s2) VPFR PFR volume (m3)
HR height of the riser (m) V total reaction volume (m3)
Hti height of i stage (m) W solid inventory in the bed (kg)
i number of stages XB average solid conversion
k kinetic pre-exponential factor (s1) z = z4 height of the riser (m)
kCSTR kinetic pre-exponential factor in CSTR reactors (s1) z1 height of dense phase (m)
kPFR kinetic pre-exponential factor in PFR reactors (s1) zi height of i stage (m)
Kbc mass transfer coefficient between bubble and cloud d diffusion coefficient (m2/s)
(s1) l gas viscosity (kg/(m s))
Kbe overall mass transfer coefficient between bubble and eb gas voidage in the bubble phase
emulsion (s1) ee gas voidage in the emulsion phase
Kce mass transfer coefficient between cloud and emulsion ef gas voidage in the whole fluidised bed
(s1) emf gas voidage at minimum fluidisation conditions
ks kinetic pre-exponential factor (m/s) es(z) solid voidage at generic height z
Lf height of the fluidised bed modelled (m) es ⁄ solid voidage at the exit of the fluidised bed
Lm height of the packed solid loading (m) esd solid voidage in the dense phase
P operative pressure (atm) esi solid voidage in the i stage
Qb total volumetric flowrate of the bubble phase (m3/s) qg gas density (kg/m3)
Qe total volumetric flowrate of the emulsion phase (m3/s) qs solid density (kg/m3)
r reaction rate (kmol/(m3 s)) r volumetric fraction of the bubble phase
R external radius of the particle (m)
r0 reaction rate (kmol/(m2 s))
rc particle radius (m)

CO2 capture include monoethanolamine (MEA) scrubbing technol- it is reduced through combustion, and the air reactor, where it is
ogy. In this process, the cooled flue gases enter into the absorber oxidised in contact with air (Fig. 1).
unit where fresh amine solvent is used to remove CO2 from the A generalised description of the overall reaction in the fuel reac-
gas stream. The spent amine solvent is regenerated in the stripper tor can be written as follows:
unit where the temperature is higher than in the absorber unit;
ð2n þ mÞMy Ox þ Cn H2m ! ð2n þ mÞM y Ox1 þ mH2 O þ nCO2 ð1Þ
CO2 is then recovered at lower pressure. MEA and other solvents
present drawback such as equipment corrosion in presence of O2 The reduced metal oxide MyOx1 is then transported to the air
and energy intensive solvent regeneration [3–5]. In addition, the reactor where it is re-oxidised to MyOx:
presence of flue gas contaminants, such as SOx and NOx, has a neg-
My Ox1 þ 1=2O2 ðairÞ ! M y Ox þ ðair : N2 þ unreacted O2 Þ ð2Þ
ative impact on solvent based process performance.
Materials such as zeolites, alumina molecular sieves and acti- Thus, the air is never mixed with the fuel avoiding NOx emis-
vated carbon are often employed for selectively adsorb CO2 in sions [12] and producing a stream of CO2 and H2O vapour; the lat-
applications for the production of H2 from syngas and in natural ter can be easily separated from the CO2 through condensation. The
gas sweetening; however, intensive adsorbent regeneration often gas stream leaving the air reactor contains nitrogen and un-reacted
implies a high energy penalty [6]. Porous membranes are also used oxygen. These gases can be released to the atmosphere with min-
to separate gases of different molecular sizes. However, since the imal negative environmental impact.
amount of gas treated is low, various operational set-ups need to Transition metals are good oxygen carrier. In particular, NiO/Ni,
be implemented in practice; those comprise multistage operation CuO/Cu, Fe2O3/Fe3O4, Mn2O3/Mn3O4 have been investigated with
or stream recycling, for example [6]. different inert support materials such as: Al2O3, TiO2, SiO2, ZrO2,
Compared to the methods described above, Chemical Looping and bentonite, to increase their reactivity, durability and fluidising
Combustion (CLC) is considered to be both pre-combustion capture properties [13]. Hossain et al. [13] discussed the different proper-
and oxy-combustion and it is potentially the technology best sui- ties obtained when different metal oxides and inert support mate-
ted for the efficient, low cost and low energy capture of CO2 from rials are mixed. The ideal oxygen carrier particles should show
flue gases [7–11]. CLC may operate, in principle, with a variety of good oxygen-carrying capacity, high oxidation and reduction reac-
fuel types, including carbonaceous fuel such as coal-derived syngas tivity, good mechanical strength, suitable particle size, density and
and natural gas. The process uses a transitional metal oxide as oxy- pore structure to achieve high fluidisability and reaction rate.
gen carrier to transfer the oxygen from the air to the fuel reactor. The hot air leaving the combustor is used to drive a steam tur-
The oxygen carrier is circulated between the fuel reactor, where bine/gas turbine combined cycle system for electricity generation.
48 R. Porrazzo et al. / Fuel 136 (2014) 46–56

Fig. 1. Circulating fluidised bed system for CLC.

Whereas the reduction reaction of the metal oxide is often endo- reactors, usually PFRs and CSTRs) combined in a fashion that could
thermic, the oxidation reaction of the metal oxide is exothermic. simulate the real hydrodynamics and/or kinetics. Following this
The heats of reaction depend on the fuel type and on the metal strategy, fluidised beds have been implemented in Aspen Plus for
oxide used as oxygen carrier. The heat integration between the a number of process set-ups (e.g. Sarvar-Amini et al. [17] modelled
two reactors can reduce the energy loss by recovering the low- a fluidised bed membrane reactor; Sotudeh-Gharebaagh et al. [18]
grade heat while producing a larger amount of high grade heat and Liu et al. [19] modelled a fluidised bed for coal combustion;
as shown in Fig. 2 [14]. Sohi et al. [20] have modelled a fluidised bed for gas natural
The overall generation of heat equals the heat of combustion. combustion).
Depending on the metal oxide utilised, the thermal energy released To the best of our knowledge, fluidised beds for CLC purpose
in the oxidation reactor is usually larger than the energy required have been modelled exclusively by using a Gibbs reactor which
by direct combustion of the fuel [14,15]. Also, the heat absorbed in is based on the minimisation of the Gibbs free energy [14,21]. Such
the reduction reactor is at low temperature and heat is released at minimisation leads to an overestimation of the conversions of the
high temperature in the oxidation reactor. These features imply species since effects such as gas by-pass and mass transfer and/or
that the combustion system can be highly efficient [15]. The free- kinetic limitations, for instance, are neglected. Those effects are of
of-water CO2, obtained after condensation of the water vapour, importance in determining the real conversion. Additionally, if
can be captured or/and used for other applications. ideal conditions are assumed, i.e. the Gibbs reactor is employed,
When information on the CLC process is needed, process simu- the real reactor size cannot be estimate.
lation such as Aspen Plus, Aspen Hysys, PRO/MAX, PRO/II are The aim of this work is to attempt to model fluidised bed reac-
implemented. Those packages are largely used for whole plant tors which mimic the real fluidised bed reactors employed in CLC.
modelling given their ability to simulate a variety of steady-state Aspen Plus is adopted by using different combinations of standard
processes ranging from single unit operation to complex processes blocks of reaction such as CSTRs and PFRs to take into account
involving many units. Such codes work with different standard different hydrodynamic regimes. Additionally, heterogeneous,
blocks that represent the main unit operations in the simulated non-catalytic reactions are incorporated in the model for the first
process (e.g. PFR and CSTR reactors, absorbers, distillation columns, time, to take into account the kinetics of reaction (shrinking-core
etc.). Mass and heat balances are within the blocks. Inlet and outlet model). An estimation of the sizes and of the optimal operating
mass and heat streams link the blocks to each other. When an conditions for the CLC reactors is obtained and comparisons are
attempt is made to simulate CLC processes with one of the process made with the Gibbs reactor, normally implemented when model-
packages, fluidised beds would need to be represented, and those ling fluidised beds in Aspen Plus [14,21].
are not employed specifically in of the packages aforementioned.
Nevertheless, an accurate representation of the process cannot 2. System description
neglect the complex hydrodynamics and kinetics happening in
the reactors. A way to solve this issue has been proposed by Jafari In this work, the most common set-up of the CLC process is
et al. [16] who employed a number of basic blocks (e.g. ideal assumed, and therefore two interconnected fluidised beds working
in different hydrodynamic regimes are considered. Specifically, the
reduction of the metal oxide is carried out in a bubbling bed while
the air reactor, where the metal oxides are oxidised, is assumed to
work in the fast fluidisation regime.

2.1. The fuel reactor

The bubbling regime is characterised by a value of the superfi-


cial gas velocity, Uo, higher than the fluidisation velocity, Umf, and
lower than the terminal velocity of the isolated particles, Ut [22].
The bubbling bed can be modelled according to the two phase the-
ory [22] where the bubble phase, with low content of solids, and
the emulsion phase, characterised by perfect mixing of gas and
solids, are considered as co-existing within the bed. According to
Davidson and Harrison [23], the emulsion phase is at minimum
Fig. 2. Heat integration scheme for CLC. fluidisation conditions and the gas in excess with respect to the
R. Porrazzo et al. / Fuel 136 (2014) 46–56 49

minimum fluidisation velocity (Uo > Umf) is transferred to the bub-


ble phase. Additional assumptions are [23]:

 The bubble diameter, db, is constant along the bed height.


 The reactor operates at isothermal conditions.
 The radial mass solid gradient within the bed is neglected.

Table 1 reports the correlations and equations adopted in this


work for the bubbling bed.

2.2. The air reactor

Fast fluidisation is characterised by the superficial velocity of


the inlet gas, Uo, greater than the terminal velocity of an isolated
particle, Ut [22]. In fast fluidisation, perfect mixing of the gas and
the solid is assumed. The solid volume fraction is assumed to Fig. 3. Trend of the solid fraction (1  e) in the riser.
remain constant in the bed radial direction, while two zones along
the bed height are identified: the dense and the lean phases; the
have been considered including CO, H2, CH4 [12,24,25]. Many
latter is divided into lower acceleration region, upper acceleration
authors have used kinetic models, based on shrinking core and
region and completely fluidised region [18,19,22] (Fig. 3). The rel-
changing grain size, to represent the chemical kinetics of the metal
ative height of those regions varies with the inlet superficial gas
oxides [13]. The best fit with experimental data is achieved with
velocity. The solid voidage of the dense phase is assumed constant
one of these two models depending on the gas reactant and the
whereas the solid voidage of the lean phase decreases along the
metal oxides considered.
bed height. The lean phase is usually referred to as the transport
To describe the reactions on the solid particles the un-reacted
disengagement height (TDH) and an exponential change in the sol-
core model is considered in this work; this is a good approximation
ids loading is assumed to describe the variation of voidage, e, along
of the real behaviour, as confirmed by a number of studies
the TDH [22]:
[13,24,26–29]. The model assumes that the reaction happens on
es  es ðzÞ az the surface separating the un-reacted solid from the reacted shell.
¼e ð3Þ
es  esd The initial reaction surface corresponds to the initial external sur-
face of the solid. The thickness of the reacted shell increases with
where es is the solid voidage at the exit of the fluidised bed, esd is the
time, determining the shrinking core of un-reacted solid [24,30].
solid voidage of the dense phase, a is the decay index. esd is a func-
The heterogeneous reaction proceeds via three steps: external
tion of inlet superficial gas velocity.
mass transfer, internal mass transfer (internal diffusion within
the particle), chemical reaction.
2.3. Reduction and oxidation kinetics Within the fuel reactor the reduction reaction is:

In this work, pure methane and NiO/Ni oxygen carrier sup- CH4 þ 4NiO ! CO2 þ 2H2 O þ 4Ni ð5Þ
ported by bentonite (80 lm diameter) are chosen as fuel and solid
While within the riser the oxidation reaction is:
reactant, respectively. Pure air is used in the air reactor to oxidise
Ni metal. This reaction has been chosen because of the large O2 þ 2Ni ! 2NiO ð6Þ
amount of kinetic data available from the literature useful to
implement and validate the model proposed. The following assumptions are made [31]:
The heterogeneous non-catalytic reaction occurring in both the
riser and fuel reactor is:  The particles are spherical.
 The external mass transfer step is fast compared to the internal
Agas þ bBsolid ! cC products : ð4Þ diffusion and reaction steps.
 The reaction is first order with respect to the concentration of
Various studies have been carried out to characterise the reduc-
the reactant gas.
tion and oxidation behaviour of metal oxides and diverse gases
 The particle volume remains constant.
 The reaction is isothermal.
Table 1
List of correlations and equations applied for the bubbling bed model. Ruy et al. [24] demonstrate that the reduction rate for NiO/Ni
Superficial gas velocity in the emulsion U mf particles supported by bentonite is controlled by the chemical
U e  emf
phase reaction while the oxidation is controlled by the internal diffusion.
Rise bubbles velocity at Uo = Umf Ubr = 0.711 * (g * db)0.5 Garcia-Labiano et al. [26,27] studied the CLC kinetics and the vari-
Rise bubbles velocity at Uo – Umf Ub = Uo  Ue + Ubr
ations in the structure of the oxygen carrier were considered
Volumetric fraction of bubble phase r ¼ UUob U e
U e together with various geometries; the changing grain size model
Fluidised bed voidage ef = r * eb + (1  r) * ee
W
was utilised. Small particles (30–70 lm) were selected to minimise
Height of the packed solid loading Lm ¼ q At ð1 em Þ
s
ð1em Þ
mass transfer limitations. The shrinking core model with the reac-
Height of fluidised bed Lf ¼ Lmð1 ef Þ
! tion being the controlling step describes well the experimental
Mass transfer coefficient between  
K bc ¼ 4:5  Ue
þ 5:85  d0:5 g 0:25 data [13]. Indeed, the oxygen carrier particles used in CLC have
bubble and cloud db 5
db4 small diameter and high internal porosity, consequently the
Mass transfer coefficient between cloud  0:5
K ce ¼ 6:77 
demf U b assumption is shown to be a reasonable one [28].
and emulsion 3
db
Data from the literature report that oxidation and reduction
Overall mass transfer coefficient 1
¼ K1ce þ K1bc
between bubble and emulsion
K be
times of the oxides selected in this study are of the same order
of magnitude. Sung and Sang [32] found a reduction conversion
50 R. Porrazzo et al. / Fuel 136 (2014) 46–56

rate of 4%/min for NiO supported by bentonite and an oxidation as plug flow whilst completely mixed conditions are assumed in
conversion rate of 12%/min in experiments with circulating fluid- the emulsion phase. The whole reactor can be divided along the
ised beds at constant temperature. Lyngfelt et al. [33] reported a axial direction into several stages, with each stage considered as
reduction conversion rate of 4%/min for pure NiO and 7%/min for made up of two parallel ideal sub-reactors: a PFR to represent
NiO supported by bentonite (60/40%) and an oxidation conversion the gas flow through the bubbles and a CSTR to represent the gas
rate of 13%/min for pure Ni and 21%/min for Ni supported by flow through the emulsion. A first order reaction with respect to
bentonite (60/40%) in fluidised beds. Garcia-Labiano et al. [26] the concentration of the gas is assumed in each sub-reactor. Mass
reported high reactivity of nichel oxygen carriers with full conver- transfer between the two sub-reactors in each stage (i.e. the PFR
sion for both reduction and oxidation reactions in the range of and CSTR) occurs at their respective exit streams [16], before enter-
20–60 s that means 300–100%/min of conversion respectively. ing the next stage, as shown in Fig. 4.
Based on these results, the same kinetic pre-exponential factor ks Aspen Plus provides calculator blocks where user defined calcu-
([=] m/s) for the oxidation and reduction is assumed; the reaction lations can be inserted; those can be written in FORTRAN or Excel
is the controlling step: can be used. Such calculators are used to link the sub-reactors in
each stage as well as each stage to the next. Specifically, the calcu-
area particles
r¼  r0 ð7Þ lator block MTr Ci (in Excel) is used to modify the outlet molar gas
reaction v olume
flow-rate from each sub-reactor by solving the mass transfer term
between bubble and emulsion phase (Eqs. (12)–(15)). Transfer
r 0 ¼ ks  C A ð8Þ
block functions (i.e. TrBi and TrEmi – see Fig. 4), insure that the
3
where CA is the molar concentration of the gas ([=] kmol/m ). streams between stages verify mass continuity for each compo-
For spherical particles: nent. In this way, the gas molar flow-rate along the bed is always
redistributed between bubble and emulsion phase. Another calcu-
area particles 4  p  r 2c  N part
¼ lator block (Feed C), written in Excel, is used to define the feed
reaction v olume reaction v olume
conditions. For a fixed value of solid inventory, W, as the inlet
4  p  r 2c  V tot  ð1  eÞ superficial velocity of the methane, Uo, varies, the volume of each
¼ ð9Þ
4
3
 p  R3  V tot  e sub-reactor varies (the volumetric fraction of the bubble phase,
r, the gas velocity of the bubbles, Ub, the volumetric flowrate of
where rc is the average particle radius at the reaction surface, R is
the bubble phase, Qb, are fixed once the superficial velocity is
the external radius of the particle, Vtot is the whole volume of the
fixed). However, a change in the solid inventory determines a var-
system and e is the voidage of the system. According to the shrink-
iation of the total height of the fluidised bed, as reported in Table 1,
ing core model [30]:
which results in a change in the volume of each sub-reactor.
1
r C ¼ R  ð1  X B Þ3 ð10Þ Aspen Plus provides only a limited number of kinetic rate
expressions; therefore, a FORTRAN code is written (i.e. Kin Ci –
where XB is the average conversion of the solid reactant. Finally the
see Fig. 4) to implement the un-reacted core model in each sub-
kinetic rate expression is given by:
reactor. Wegstein convergence solver [34] is used for solving the
6  ð1  eÞ  ks  C A  ð1  X B Þ3
2
2
mass balance on the streams from and to the sub-reactors in each
r¼ ¼ k  C A  ð1  X B Þ3 ð11Þ stage. Mass balance equations for each component in each stage
dp  e
are reported in Table 2.
where k is the kinetic pre-exponential factor ([=] s1). Considering the mass transfer between the emulsion and the
bubble phase, the values of methane concentrations at the begin-
3. Aspen Plus implementation ning of the i + 1 stage are defined as:
Hti
3.1. Bubbling bed C CH4bðiþ1Þ ¼ C CH4bi  K be  ðC CH4bi  C CH4ei Þ  ð12Þ
Ub
The hydrodynamics of bubbling beds is rather complex (e.g.,
Hti  r 
[22,23]). In this work, the gas flow through the bubbles is treated C CH4eðiþ1Þ ¼ C CH4ei þ K be  ðC CH4bi  C CH4ei Þ   ð13Þ
Ue 1r

Table 2
Mass balance equations for each component in each stage of the bubbling bed model.

Bubble phase
R zi
C CH4bði1Þ  U b  Ab  C CH4bi  U b  Ab  K be  ðC CH4bi  C CH4ei Þ  V bi  Ab  eb  zi1 r CH4i dz ¼ 0
R zi
F NiObði1Þ  F NiObi  4  Ab  eb  zi1 r CH4i dz ¼ 0
R zi
F CO2bði1Þ  F CO2bi þ Ab  eb  zi1 r CH4i dz ¼ 0
R zi
F H2Obði1Þ  F H2Obi þ 2  Ab  eb  zi1 r CH4i dz ¼ 0
R zi
F Nibði1Þ  F Nibi þ 4  Ab  eb  zi1 r CH4i dz ¼ 0

Emulsion phase
 d 
C CH4eði1Þ  U e  Ae  C CH4ei  U e  Ae þ K be  ðC CH4bi  C CH4ei Þ  V ei  1d  r CH4  V CSTRi ¼ 0
FNiOe(i1)  FNiOei  4 * rCH4 * VCSTRi = 0
FCO2e(i1)  FCO2ei + rCH4 * VCSTRi = 0
FH2Oe(i1)  FH2Oei + 2 * rCH4 * VCSTRi = 0
FNie(i1)  FNiei + 4 * rCH4 * VCSTRi = 0
Bubble volume Vb Vb = V * r
Emulsion volume Ve Ve = V * (1  r)
PRF volume VPFR VPFR = Vb * eb
CSTR volume VCSTR VCSTR = Ve * ee
R. Porrazzo et al. / Fuel 136 (2014) 46–56 51

Fig. 4. Aspen Plus bubbling bed scheme.

Table 3
where CCH4bi and CCH4ei are the methane concentrations at the exit Parameters used for the simulation of the bubbling bed.

of the i stage for the bubble and the emulsion phase, respectively; Parameter Value Units Source
Kbe is the overall mass transfer coefficient between the bubble T 750 °C Ref. [24]
and the emulsion phase; Ub and Ue are the bubble and the emulsion P 1 atm Assumed
superficial gas velocity, respectively; Hti is the height of i stage. dp 8.00E05 m Ref. [24]
A different expression is applied if the concentrations of the qg 0.191 kg/m3 Aspen Plus database
qs 2489 kg/m3 Aspen Plus database
component at i + 1 stage become negative or if the mass driving
l 0.000027 kg/(m s) Aspen Plus database
force principle is not satisfied: DBB 1 m Assumed
  At 0.78 m2 Calculated
1 Q Umf 0.0096 m/s Calculated
C CH4bðiþ1Þ ¼  C CH4bi þ C CH4ei  e ð14Þ
2 Qb Ue 0.019 m/s Calculated
db 0.03 m Ref. [25]
  Ut 0.26 m/s Calculated
1 Q Ubr 0.38 m/s Calculated
C CH4eðiþ1Þ ¼  C CH4ei þ C CH4bi  b ð15Þ Ub 0.46 m/s Calculated
2 Qe
Uo 0.1 m/s Assumed
Fs 0.277 kg/s Assumed
where Qb and Qe are the total volumetric flowrates of the bubble
NiO inlet 98 % w/w Assumed
and the emulsion phases, respectively. The distribution of the reac- Ni inlet 2 % w/w Calculated
tant gas between the two phases affects largely the reaction; whilst emf 0.5 – Assumed
the methane mass transfer is explicitly considered (see Eqs. (12)– eb 0.9 – Assumed
(15)), the mass transfer of the products (CO2 and H2O) between Lm 1 m Assumed
em 0.45 – Assumed
the two phases is assumed to be negligible since the kinetics is
r 0.18 – Calculated
assumed to be unaffected by the products. ef 0.57 – Calculated
An increase in the number of stages results in an increase in the Lf 1.28 m Calculated
amount of gas transferred from the bubble phase to the emulsion d 6.50E05 m2/s Aspen Plus database
Kbc 9.57 s1 Calculated
phase and thus available for the reaction. The effect of the number
Kce 5 s1 Calculated
of stages on the final gas conversion varies depending on the Kbe 3.31 s1 Calculated
hydrodynamic and kinetic variables, i.e. on the ratio between the ks 4.41E04 m/s Ref. [24]
inlet superficial gas velocity Uo and the inlet superficial gas velocity kCSTR 33.06 s1 Calculated
at minimum fluidisation condition Umf, and the ratio between the kPFR 3.67 s1 Calculated

kinetic coefficient k and the mass transfer coefficient Kbe. When


Uo  Umf, since it is assumed that the bubble phase takes the
excess of gas, with respect to the emulsion phase, the volumetric gas is present in the emulsion phase; in this case the gas conver-
fraction of the bubble phase, r, is high; additionally, if the number sion is slightly affected by the mass transfer and by the number
of stages increases, the mass transfer between the two phases also of stages: the effect of the by-pass of gas in the bubble phase is
increases and so the conversion (more gas available for the reac- not very relevant since the gas in the emulsion phase is not largely
tion). When Uo  Umf, the volumetric bubble fraction r is low com- consumed.
pared to the emulsion volumetric fraction, and the amount of gas The right number of stages to model the system depends on
that by-passes the emulsion phase is small; in this case, the pres- its kinetics and hydrodynamics and the details of the choice made
ence of the bubbles is less relevant and the mass transfer between will be discussed below. Experimental data taken from the liter-
the two phases is low. In this situation, the number of stages ature are used to validate the multi-stages model [13,14]. In
affects slightly the gas conversion. When k  Kbe, the mass transfer Table 3 the values of the parameters used for the simulation of
between the two phases is the controlling step and most of the the bubbling bed are reported. The solid fed into the system is
unreacted gas remains in the bubble phase while the gas in the in stoichiometric quantity; in the air reactor, the solid reacts up
emulsion phase is quickly consumed; an increase in the number to the given conversion (almost unity); the oxidised solid is then
of stages allows a for large amount of fresh gas coming from the sent into the fuel reactor where all the methane reacts and the
bubbles to react with the solid particles. When k  Kbe the kinetic final conversion is dictated by the oxidised solid circulating in
term becomes the controlling step and a large amount of unreacted the system.
52 R. Porrazzo et al. / Fuel 136 (2014) 46–56

3.2. Riser implementation Table 4


Mass balance equations for each component in each CSTR of the fast fluidisation
model.
The different amount of solid in the different regions of the riser
affects the reaction. Since the kinetic rate changes depending on FO2(i1)  FO2i  rO2i * VCSTRi = 0
the solid voidage (Eq. (11)), the riser is split into a number of CSTRs FNi(i1)  FNii  2 * rO2i * VCSTRi = 0
FNiO(i1)  FNiOi + 2 * rO2i * VCSTRi = 0
in series [18,19]. In the case under exam, the system is split into
four CSTRs: one CSTR is assumed to represent the dense phase
and three CSTRs are assumed to mimic the lean phase character-
ised by three different mean voidages calculated in the following
way: Table 5
  List of the values of parameters used for the simulation of the riser.
esd  es
esi ¼ es   ðeazi  eazi1 Þ ð16Þ Parameter Value Units Source
a  ðzi  zi1 Þ
T 750 °C Assumed
with i = 2, 3, 4. z1 is the height of the dense phase which, summed of P 1 atm Assumed
all to other zones, up to z4, gives the height of the whole riser. The dp 8.00E05 m Assumed
relationship between a and Uo plotted by Kunii and Levenspiel [22] qg 0.23889 kg/m3 Aspen Plus database
qs 6033 kg/m3 Aspen Plus database
shows that the product of a times Uo is a constant and its value is
l 4.7E05 kg/(m s) Aspen Plus database
between 4 and 12. DR 0.5–1 m Assumed
The previous system of three equations (Eq. (16)) is solved with HR 3.5–4.5 m Assumed
the following constrains: At – m2 Calculated
Umf 0.013 m/s Calculated
z2  z1 ¼ z3  z2 ¼ z4  z3 ð17Þ Ut 0.35 m/s Calculated
Uo – m/s Calculated
z 4 ¼ z 1 þ 3  Dz ð18Þ Fs 0.22 kg/s Exit fuel reactor
NiO 5 % w/w Exit fuel reactor
And the constrain of the solid mass balance: Ni 95 % w/w Exit fuel reactor
Lm 0.25–1 m Assumed
X
4 esd1 – – Ref. [22]
Lm  ð1  em Þ ¼ z1  esd þ z  esi ð19Þ esd2 – – Calculated
i¼2 esd3 – – Calculated
esd4 – – Calculated
where Lm * (1  em) is the height of the packed solid loading equal to ks 4.41E04 m/s Assumed
W/(qs * At). W is the inventory of solid in the system, qs is the solid
density and At is the area of the column.
Most of the reaction occurs in the dense zone because of the
large quantity of reactant solid present. A calculator block (Feed
C) written in Excel is used to define the operating condition of
the feed and the volume of each sub-reactor. The calculator block,
by solving Eqs. (16)–(19), determines the height of the column and
the volume of each CSTR for an assumed value of the solid inven-
tory, diameter of the column, and inlet superficial gas velocity Uo. A
FORTRAN code is written into the calculator blocks (Kin Ci) to mod-
ify in each CSTR the kinetic rate expression, as reported in the case
of the bubbling bed (Section 3.1). Broyden convergence solver [34]
is used for solving the mass balance in each CSTR. In Fig. 5 the
Aspen Plus scheme for the fast fluidisation regime is shown.
Mass balance equations for each component in each CSTR reac-
tor are reported in Table 4. In Table 5 the values of parameters used
for the simulation of riser are reported.

4. Results and discussion

4.1. Bubbling bed

In Fig. 6 the conversion of methane is reported as a function of Fig. 6. Conversion of gas vs number of stages n at different k keeping Uo and Fs
the number of stages. The inlet gas superficial velocity, Uo, and the constants.

Fig. 5. Aspen Plus fast fluidisation scheme.


R. Porrazzo et al. / Fuel 136 (2014) 46–56 53

Fig. 7. Conversions of both gas and solid vs Uo at n = 5 keeping Fs constant. Fig. 9. Conversion of gas vs solid inventory W at n = 5.

Fig. 8. Conversions of both gas and solid vs Fs at n = 5 keeping Uo constant.

solid flowrate, Fs, are kept constant. The ratio between the mass
transfer coefficient Kbe and the pre-exponential factor k is varied
with Kbe kept constant.
As mentioned previously, a k value lower than Kbe implies that Fig. 10. Conversions of both gas and solid vs riser diameter DR.

the kinetics is the controlling step and therefore the contribution of


gas by-pass to the overall gas conversion is lower. In this case, the
overall gas conversion is not affected greatly by the number of  The gas residence time is long enough to expect the methane
stages, therefore the system can be modelled with a fewer stages. conversion higher than 90%; this condition is achieved with 5
Conversely, a higher value of k implies the mass transfer to be the stages and the conversion of both gas and solid reaches a
controlling step: consequently, a good gas redistribution between plateau.
the bubble and the emulsion phase is relevant. In this case, the var-  The gas residence time from the literatures [13,14], for methane
iation of the overall gas conversion with the number of stages is conversion higher than 90%, is found to be lower than the one
larger than the conversion obtained in the case of kinetic control. calculated; consequently, to reduce the gas residence time,
After 5 stages, the system reaches a plateau and therefore the the height of the fluidised bed is reduced (and thus the solid
multi-stage model assumes 5 stages as the maximum number of inventory – see Table 1). (As it can be seen in Fig. 9, the same
stages. In general, the increase in the number of stages determines conversion will be obtained, since for W > 350 kg, the conver-
the increase in the conversion for both gas and solid and the con- sion is practically constant.)
version values could overpredict the experimental data since the
hydrodynamics of the emulsion phase moves from well mixed to In Fig. 7 the conversions of both methane and solid, when Uo
plug flow. In the case analysed here, k is taken equal to 33 s1. varies and the inlet solid mass flow-rate is kept constant are
Experimental data taken from the literature are considered to jus- reported. A comparison with the Gibbs reactor model is presented
tify the choice made [13,14]. A number of authors have found that in Fig. 7.
the methane conversion, using NiO/Ni supported by bentonite or The increase in Uo determines a higher by-pass of the gas in the
Al2O3 in a range of temperature of 1000–1400 °K, is more than bubble phase with a consequent decrease in the gas conversion.
90% with gas residence time in the range of 1–5 s [13,14]. The When Uo decreases the reactant gas becomes the limiting reactant
gas residence time in the present system is about 13 s. This condi- and thus the solid conversion decreases. For a value of the inlet gas
tion leads to two considerations: superficial velocity higher than 0.1 m/s the solid becomes the
54 R. Porrazzo et al. / Fuel 136 (2014) 46–56

Fig. 12. Conversions of both gas and solid vs height of the packed solid loading Lm.
Fig. 11. Conversions of both gas and solid vs riser height HR.

limiting reactant and the calculation is stopped. The sharply


increase of solid conversion depends on the high kinetic rate (k is
equal to 33 s1) of NiO reduction. Runs with lower kinetic rate
are carried out and they show a slow increase of the solid conver-
sion when Uo increases; this is a consequence of the fact that the
decrease of both gas residence time and gas–solid contact time
becomes more relevant. Finally, the slope of the solid conversion
vs the inlet superficial gas velocity Uo is a function of the kinetic
rate.
Considering the comparison between the model proposed here
and the largely used Gibbs reactor model, the Gibbs reactor
assumes that the free energy of the reaction is minimum. Thus,
the hydrodynamics is not considered and no information is avail-
able on the size of the reactors. The comparison between the two
models shows that the results are close when the inlet superficial
gas velocity is close to Umf (equal to 0.01 m/s): in this case, the vol-
umetric fraction in the bubble phase is small and thus its influence
on the overall gas conversion can be neglected. Most of the gas is
consumed in the CSTR reactors: the assumption that the reaction
kinetics is fast, allows for a nearly complete methane conversion.
When Uo increases, the volumetric fraction of the bubble phase
increases and thus gas bypass is observed; consequently, the dif-
ference between the present model and the Gibbs reactor becomes Fig. 13. Conversions of both gas and solid vs molar stoichiometric ratio Fair/FCH4.

more pronounced. In this case, a Gibbs reactor gives higher conver-


sions than the multistage system implemented here. Consequently, the new value of 360 kg is chosen and that corre-
In Fig. 8 the conversions of both methane and solid vs Fs, at con- sponds to a value of Lm of 33 cm. This corresponds to a reaction
stant Uo and n equal to 5, are shown. It can be seen that when Fs time of 21 min and a solid conversion rate of 4.6%/min: such value
increases the solid conversion decreases for a constant value of is very close to the experimental data reported by Lyngfelt et al.
Uo. This is a consequence of the fact that for a value of Fs higher [33] (7%/min of reduction reaction rate with NiO supported 60/
than 0.277 kg/s the gas becomes the limiting reactant. The solid 40 and 4%/min of reduction reaction rate with pure NiO) for NiO/
residence time for the reduction reaction of the NiO expressed as Ni supported reacting with methane. Based on measured kinetics
ratio between the whole mass of solid charged into the system and assuming complete gas–solid contact, a very small solid
W and the stoichiometric solid flowrate [33] is 64 min: that corre- inventory of around 10–20 kg/MW would be sufficient to reach full
sponds to a solid conversion rate of 1.55%/min. This reaction time conversion in the fuel reactor, or equivalently, to reach the equilib-
is quite large. A sensitivity analysys is necessary to know the right rium conversion [28]. However, in bubbling fluidised beds, the
loading of solid to reach more than 90% in the gas conversion. In gas–solids contact is not complete, since gas by-pass is induced
Fig. 9 the relationship between the solid inventory and the gas con- by the presence of the bubbles. Consequently, predictions based
version is shown. on the kinetics and assuming complete gas–solids contact would
Fig. 9 shows that in the interval 360–1060 kg of solid inventory underpredict the solid inventory, and much larger solid inventory
the gas conversion is unchanged. Thus, the same degree of gas would be needed, around 500 kg/MW [29,35]. In this case, the
conversion can be achieved by choosing a lower solid inventory. plant generates 0.75 MW, thus the solid inventory from the
R. Porrazzo et al. / Fuel 136 (2014) 46–56 55

Table 6 the same temperature of the fuel reactor are assumed. Higher tem-
Optimised values of the variables that affect the riser performance. perature and higher values of Ks will lead to higher conversion as
Uo (m/s) DR (m) HR (m) Lm (m) ks (m/s) Fair/FCH4 founded by Lyngfelt et al. [33] (21%/min) and Sung and Sang [32]
1.79 0.8 3.5 0.25 4.41E04 1.25 (12%/min). The value of the inlet superficial gas velocity, Uo,
obtained is in good agreement with data from the literature (about
100 Umf) [14]. Volumetric gas flow in the air reactor is approxi-
mately 10 times larger than that of CH4 (11 times in this case) as
Table 7 mentioned by Abad et al. [28] and Johansson et al. [36].
Heat duty calculation of the CLC system.
4.3. Net heat duty
Heat duty
Bubbling bed (T = 750 °C) Riser (T = 750 °C)
Finally, a check of the net heat duty of the whole process is car-
Cstr1 BB 24.06 kW pfr1 BB 2.69 kW Cstr1 415.58 kW ried out. It is expected that the net heat duty of the process
Cstr2 BB 58.87 kW pfr2 BB 0 kW Cstr2 296.65 kW
matches the net heat duty produced by the direct combustion of
Cstr3 BB 30.69 kW pfr3 BB 0 kW Cstr3 78.07 kW
Cstr4 BB 16.49 kW pfr4 BB 0 kW Cstr4 45.28 kW methane with air [14,15]. The direct combustion of methane is
Cstr5 BB 8.97 kW pfr5 BB 0 kW implemented in Aspen Plus using a stoichiometric reactor at the
Tot BB 141.76 kW Tot Riser 835.57 kW same temperature of the two fluidised beds (750 °C), with 92.5%
of methane molar conversion as obtained in the fuel reactor. The
simulated results show that the net heat duty of the process is
693.81 kWatts that is equal to the net heat duty produced by
literature should be around 375 kg: this value is close to the oper-
methane direct combustion (Table 7). In the riser, the heteroge-
ating value of 360 kg.
neous exothermic reaction occurs mainly in the dense phase and
the heat produced decreases along the bed height as a consequence
4.2. Riser of the decreasing of the reaction due to a decrease of solid particles.
In the bubbling bed reactor, the heat consumed by the endother-
The inlet solid flowrate in the riser (air reactor) must match the mic reaction decreases along the bed height; this is a consequence
rate circulating in the bubbling bed fuel reactor. Usually a cyclone of the decrease in methane concentration resulting in the reduc-
between the two beds allows for the separation between gas and tion of the reaction; this trend is observed from CSRT2 up to CSRT5.
solid products from the air reactor. The solid particles from the CSTR1 takes care of the split of the feed between bubble and emul-
bubbling bed return to the air reactor by gravity. sion phase and it is introduced to initialise the sequence of the
In the air reactor the variables that can be changed, apart from stages. Finally, the CLC power plant is assumed to work at atmo-
operative temperature and pressure, are: diameter DR, height HR, spheric pressure, with the fuel reactor usually working in adiabatic
solid inventory and molar stoichiometric ratio between air and fuel conditions and with the circulating solid flow-rate from the air
reactants. A sensitivity analysis is carried out to understand how reactor sustaining the endothermic reaction [14,15]; the riser usu-
the variables taken into account influence the conversions of both ally works at temperatures higher than those in the fuel reactor
gas and solid. The sensitivity analysis gives information on the val- and the difference in temperature depends on the amount of solid
ues of the variables that optimise the process. In Fig. 10 the conver- flow-rate circulating between the two reactors. The extra energy
sions of both gas and solid, as a function of the riser diameter, are produced by the exothermic reaction in the riser is employed in
shown. The increase in the diameter of the riser, DR determines a a steam turbine cycle for electricity production.
decrease in Uo (at constant inlet molar air flowrate), which implies
a higher gas residence time, higher solid–gas contact time and 5. Conclusions
higher conversions. In Fig. 11 the conversions of both gas and solid
as function of the height of the riser HR are shown: In this work, a circulating chemical looping combustion process
An increase in the height of the riser, HR, corresponds to a higher is studied. The system is characterised by a fuel reactor operating
residence time for the gas and higher conversions. In Fig. 12 the con- in the bubbling regime, and a circulating bed air reactor operating
versions of both gas and solid, as a function of the solid inventory, in the fast fluidisation regime. The reactions taking place are
expressed as height of packed solid loading Lm (Table 1), are shown. non-catalytic heterogeneous reactions and the shrinking core model
The increase in the height of packed solid loading determines a is applied to describe the kinetic mechanism. The Davidson and
higher kinetic rate reaction as a function of solid voidage (Eq. (11)) Harrison model [23] is applied to describe the hydrodynamics of
and therefore higher conversions. In Fig. 13 the conversions of both the bubbling bed and bubble phase and emulsion phase are consid-
gas and solid, as a function of the molar stoichiometric ratio ered. The riser is described considering only one phase with solid
between air and methane, are reported. It can be noted that the particles concentration varying along the bed height. Air and meth-
increase of the molar stoichiometric ratio between air and meth- ane are considered as inlet gas reactants for the riser and the fuel
ane determines an increase of the solid conversion and a decrease reactor respectively. NiO/Ni is used as oxygen carrier between
of gas conversion because of the excess of gas with respect to the the two fluidised beds which undergo red-ox reactions. The whole
stoichiometric amount. system is modelled in Aspen Plus where modifications are made to
Analysis of Figs. 10–13 leads to the optimal choice of the values consider the hydrodynamics and the kinetics of the system in a
of the variables that affect the riser performance and those are novel way. The results obtained show that:
reported in Table 6.
The oxygen molar conversion is about 73% and the nickel molar  The model allows for including hydrodynamic and kinetic
conversion is about 96%. The data reported in Table 6 allow for the mechanisms although its complexity is lower than CFD models
solid flowrates in the two reactors to match. In the riser the solid implementing more rigorous description of the system at the
inventory is about 417 kg, the solid oxidation reaction time is particle level (e.g. two-fluid models).
about 31 min and the oxidation solid conversion rate is about  The model gives a good estimation of the main variables
4%/min. The solid oxidation conversion rate is close to the reduc- involved, namely diameter and height of the two beds, solid
tion one because the same kinetic pre-exponential factor Ks and inventory, molar stoichiometric ratio between air and fuel.
56 R. Porrazzo et al. / Fuel 136 (2014) 46–56

 A comparison with a fluidised bed modelled using the Gibbs [13] Hossain MM, de Lasa HI. Chemical-looping combustion (CLC) for inherent CO2
separations – a review. Chem Eng Sci 2008;63:4433–51.
reactor is carried out showing differences in terms of gas/solid
[14] Fan Liang-Shih, John Easton C. Chemical looping systems for fossil energy
conversion since the gas by-pass through the bubble phase is conversions. Wiley; 2010.
neglected in a Gibbs reactor model; consequently, the model [15] Hatanaka T, Matsuda S, Hatano H. A new-concept gas–solid combustion
proposed here is more accurate (realistic) than the Gibbs one. system ‘‘MERIT’’ for high combustion efficiency and low emissions. In:
Proceedings of intersociety energy conversion engineering conference, vol.
 The model makes it possible to estimate the thermal efficiency 30; 1997. p. 944–8.
of the process and to undertake an economic analysis. [16] Jafari R, Sotudeh-Gharebagh R, Mostoufi N. Modular simulation of fluidized
reactors. Chem Eng Technol 2004;27:2.
[17] Sarvar-Amini A, Sotudeh-Gharebagh R, Bashiri H, Mostoufi N, Haghtalab A.
As a concluding remark, more complicated reactor set-ups, Sequential simulation of a fluidized bed membrane reactor for the steam
could in principle be modelled by implementing the appropriate methane reforming using ASPEN PLUS. Energy Fuels 2007;21:3593–8.
combination of sub-reactors. [18] Sotudeh-Gharebaagh R, Legros R, Chaouki J, Paris J. Simulation of circulating
fluidized bed reactors using Aspen Plus. Fuel 1998;77(4):327–37.
[19] Liu B, Yang X, Song W, Lin W. Process simulation of formation and emission of
Acknowledgement NO and N2O during coal decoupling combustion in a circulating fluidized bed
combustor using Aspen Plus. Chem Eng Sci 2012;71:375–91.
[20] Sohi AH, Eslami A, Sheikhi A, Sotudeh-Gharebagh R. Sequential-based process
The Engineering and Physical Sciences Research Council modeling of natural gas combustion in a fluidized bed reactor. Energy Fuels
(EPSRC) support is acknowledged under the Grant No. EP/ 2012;26:2058–67.
F034482/1. [21] Li F, Kim HR, Sridhar D, Wang F, Zeng L, Chen J, et al. Syngas chemical looping
gasification process: oxygen carrier particle selection and performance. Energy
Fuels 2009;23:4182–9.
References [22] Kunii D, Levenspiel O. Fluidization engineering. 2nd ed. Boston: Butterworth-
Heinmen; 1991.
[1] Bryant EA. Climate process and change. Cambridge (UK): Cambridge University [23] Davidson JF, Harrison D. Fluidized particles. Cambridge University Press; 1963.
Press; 1997. [24] Ryu HJ, Bae DH, Han KH, Lee SY, Jin GT, Choi JH. Oxidation and reduction
[2] National Council for Science and the Environmental. In: Blockstein DE, characteristics of oxygen carrier particles and reaction kinetics by unreacted
Shockley MA, editors. Energy for a sustainable and secure future: a report of core model. Korean J Chem Eng 2001;18:831–7.
the sixth national conference on science, policy and the environment, [25] Mattisson T, Lyngfelt A, Cho P. The use of iron oxide as an oxygen carrier in
Washington, DC; 2006. chemical-looping combustion of methane with inherent separation of CO2.
[3] Abu-Zhara MRM, Schneiders LHJ, Niederer JPM, Feron PHM, Versteeg GF. CO2 Fuel 2001;80:1953–62.
capture from power plants: Part II. A parametric study of the technical [26] Garcia-Labiano F, de Diego LF, Adanez J, Abad A, Gayan P. Temperature
performance based on mono-ethanolamine. Int J Greenhouse Gas Control variation in the oxygen carrier particles during their reduction and oxidation
2007;1(1):135–42. in a chemical-looping combustion system. Chem Eng Sci 2005;60:851–62.
[4] Abu-Zhara MRM, Schneiders LHJ, Niederer JPM, Feron PHM, Versteeg GF. CO2 [27] Garcia-Labiano F, Adanez J, de Diego LF, Gayan P, Abad A. Effect of pressure on
capture from power plants: Part I. A parametric study of the technical the behavior of copper-, iron-, and nickel-based oxygen carrier for chemical-
performance based on mono-ethanolamine. Int J Greenhouse Gas Control looping combustion. Energy Fuel 2006;20:26–33.
2007;1(1):37–46. [28] Abad A, Adánez J, García-Labiano F, De Diego LF, Gayán P, Celaya J. Mapping of
[5] Singh D, Croiset E, Douglas PL, Douglas MA. Techno-economic study of CO2 the range of operational conditions for Cu-, Fe-, and Ni-based oxygen carriers
capture from an existing coal-fired power plant: MEA scrubbing vs. O2/CO2 in chemical looping combustion. Chem Eng Sci 2007;62:533–49.
recycle combustion. Energy Convers Manage 2003;44(19):3073–91. [29] Abad A, Adánez J, Dueso C, García-Labiano F, De Diego LF, Gayán P. Modeling of
[6] Herzog HJ. The economics of CO2 capture. In: Proceedings of the fourth the chemical looping combustion of methane using a Cu-based oxygen-carrier.
international conference of greenhouse gas control technologies. London: Combust Flame 2010;157:602–15.
Pergamon Press; 1999. p. 101–6. [30] Levenspiel O. Chemical reaction engineering, vol. 83. New York: Wiley; 1965.
[7] Ritcher H, Knoche K. Reversibility of combustion process. ACS Symp Ser [31] Marban G, Garcia-Calzada M, Fuertes AB. Kinetics of oxidation of CaS particles
1983;235:71–85. in the regime of low SO2 release. Chem Eng Sci 1999;54:77.
[8] Ishida M, Zheng D, Akehata T. Evaluation of a chemical-looping combustion [32] Sung RS, Sang DK. Chemical-looping combustion with NiO and Fe2O3 in
power-generation system by graphic exergy analysis. Energy 1987;12:147–54. thermobalance and circulating fluidized bed reactor with double loops. Ind
[9] Wolf J, Anheden M, Yan J. Comparison of nickel- and iron-based oxygen Eng Chem Res 2006;45:2689–96.
carriers in chemical looping combustion for CO2 capture in power generation. [33] Lyngfelt A, Leckner B, Mattisson T. A fluidized-bed combustion process with
Fuel 2005;84:993–1006. inherent CO2 separation; application of chemical-looping combustion. Chem
[10] Kerr HR. Capture and separation technologies gaps and priority research Eng Sci 2001;56:3101–13.
needs. In: Thomas D, Benson S, editors. Carbon dioxide capture for storage in [34] Gupta SK. Numerical method for engineers. New Age International; 1995.
deep geologic formations—results from the CO2 capture project, vol. 1. Oxford [35] Lyngfelt A. Oxygen carriers for chemical looping combustion – 4000 h of
(UK): Elsevier Ltd.; 2005 [chapter 38]. operational experience. Oil Gas Sci Technol 2011;66(2):161–72.
[11] IPCC special report on carbon dioxide capture and storage; 2005. [36] Johansson E, Mattisson T, Lyngfelt A, Thunman H. A 300 W laboratory reactor
[12] Ishida M, Jin H. Novel chemical-looping combustor without NOx formation. Ind system for chemical looping combustion with particle circulation. Fuel
Eng Chem Res 1996;35:2469–72. 2006;85:1428–38.

Das könnte Ihnen auch gefallen