Sie sind auf Seite 1von 17

Chemical Engineering Science 104 (2013) 233249

Contents lists available at ScienceDirect

Chemical Engineering Science


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

Heterogeneous modeling of chemical-looping combustion.


Part 1: Reactor model
Lu Han, Zhiquan Zhou, George M. Bollas n
Department of Chemical, Materials & Biomolecular Engineering, University of Connecticut, Storrs, CT 06269, USA

H I G H L I G H T S

G R A P H I C A L

A B S T R A C T

 A model for chemical-looping combustion in xed-beds is presented


and validated.
 The importance of transport contributions on kinetic rates is evaluated.
 Non-ideal ow patterns are modeled.
 Temperature uctuations are predicted for different experiments.
 An optimal oxygen carrier is proposed using the model.

art ic l e i nf o

a b s t r a c t

Article history:
Received 20 June 2013
Received in revised form
19 August 2013
Accepted 2 September 2013
Available online 14 September 2013

Heterogeneous modeling is used to investigate the signicance of intraparticle diffusion, temperature


uctuations, and dispersion on predicting xed-bed chemical-looping combustion (CLC) reactors. The
model is validated against chemical-looping reduction experiments from the literature that utilize
methane as fuel and a Ni-based oxygen carrier. The effect of particle size on CLC reduction reactions is
studied by examination of the WeiszPrater and Mears criteria for internal and external mass transfer
limitations. Homogeneous and heterogeneous models are applied for two particle sizes and it is shown
that diffusion limitations are signicant for the larger particle. Furthermore, non-ideal ow patterns are
evaluated, where superior predictions are obtained by considering the effects of dispersion and
convection. Radial dispersion in each experiment is analyzed in detail and the predicted temperature
uctuations inside the reactor and particle are presented. Lastly, the heterogeneous model is applied to
design a particle that maximizes chemical-looping efciency, measured as fuel conversion and product
selectivity. Enhancements to the overall reaction rate and temperature prole are observed for the
optimal particle size as a result of reducing the intraparticle resistance.
& 2013 Elsevier Ltd. All rights reserved.

Keywords:
Chemical-looping
Heterogeneous modeling
Oxygen carrier
Diffusion

1. Introduction
The reliance of burning fossil fuels to meet the world's increasing energy demand has contributed to the dramatic rise in global
carbon dioxide (CO2) concentration from about 280 ppm during the
pre-industrial era to 379 ppm in 2005, the highest in over 450,000
n
Correspondence to: Department of Chemical & Biomolecular Engineering,
University of Connecticut, 191 Auditorium Road Unit 3222, Storrs,
CT 06269-3222, USA. Tel.: 1 860 486 4602.
E-mail address: george.bollas@uconn.edu (G.M. Bollas).

0009-2509/$ - see front matter & 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ces.2013.09.021

years (Halmann and Steinberg, 2000; Mattisson et al., 2008). The


widely accepted impact of increasing CO2 concentration in the
atmosphere is climate change associated with global warming
(IPCC, 2008, 2007a). There is evidence suggesting that global
average surface temperature and sea level are on the rise, and in
fact, the last twelve years have been the warmest recorded years
since the beginning of instrumentally recorded temperature (IPCC,
2007b). Therefore, it is necessary to reduce greenhouse gas emissions to mitigate this impact as much as possible.
One approach to minimize CO2 emissions is to increase use of
renewable energy resources, such as biofuel, geothermal, wind

234

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

and solar. However, these renewable technologies are not


currently able to meet the ever increasing energy demand. As
dependency of fossil fuels will likely to continue in the near future,
carbon capture and sequestration (CCS) technologies are the only
viable options to reduce net CO2 emissions. In CCS, CO2 is
recovered from ue gas streams and stored in geological formations, such as aquifers or depleted gas and oil elds. The most
severe economic and energetic cost for CCS is related to the CO2
capture process (IPCC, 2005). It has been estimated that the
reduction in efciency of power process due to separation of
carbon dioxide is 810% for coal and natural gas (Finkenrath,
2011). Implementation of CCS is feasible on large stationary point
sources of CO2, and in the past few years, extensive research has
been focused on obtaining ue gas streams with high CO2
concentrations (490%) in the power production sector, which is
responsible for a third of all anthropogenic CO2 emissions (Aaron
and Tsouris, 2005).
Chemical-looping combustion (CLC) has been identied as a
promising CO2 capture technology that combines power production with CO2 capture in a single step with high efciency and low
cost (IPCC, 2005; Kerr, 2005). CLC was initially proposed to
increase thermal efciency in power generation stations, but later
was identied as having inherent advantages for CO2 separation
with minimum energy losses (Ishida et al., 1987; Richter and K.K.F.,
1983). The main difference between CLC and conventional combustion technologies is that direct contact between air and fuel is
circumvented, so that pure CO2 can be obtained without nitrogen
dilution. CLC can be classied as an oxy-fuel process, but more
adequately placed in a separate category: un-mixed combustion
(Linderholm et al., 2009). In CLC, oxygen contacts the gaseous fuel
in the form of an intermediate oxygen carrier, which is usually a
metal/metal oxide (MyOx/MyOx-1) that is alternatively oxidized
and reduced. Chemical-looping reforming (CLR) utilizes the same
general concepts as CLC, where the only difference is that the
desired product is hydrogen not heat (Fig. 1).
The CLC process is typically implemented as two interconnected uidized bed reactors, an Oxidizer and Reducer, with a
solid oxygen carrier circulating between the reactors (Johansson
et al., 2006a; Kronberger et al., 2005, 2004; Lyngfelt et al., 2001).
Complete combustion of the fuel by a metal oxide (MyOx) in the
reducer produces CO2 and water vapor, which upon condensation
produces a pure stream of CO2 ready for sequestration. The metal
oxide (MyOx) becomes reduced (MyOx  1) and is transported to the
oxidizer where it is regenerated and then cycled to the Reducer for
continued use. The outlet gas stream of the oxidizer contains
nitrogen and unreacted oxygen which can be released to the
atmosphere without negative environmental impact. In comparison with conventional combustion, CLC offers the advantage of
hydrocarbon oxidation in a nitrogen-free environment, eliminating the need of an additional energy intensive separation of CO2
from other combustion products. In addition, CLC minimizes NOx
formation because the fuel is burned in a reducing environment, in
the absence of nitrogen and at comparatively low temperatures

(o1200 1C) without ame (Hossain and De Lasa, 2008; Ishida and
Jin, 1996).
Research in the CL process has been largely focused on the
development and testing of suitable oxygen carriers. The desired
characteristics include high rates of oxidation and reduction, high
melting point temperature, low tendency of attrition and fragmentation, low toxicity, low cost of raw materials and production,
and high combustion efciency (Hossain and De Lasa, 2008).
Oxygen carriers are formulated by depositing an active component, can be Cu, Fe, Ni, or Mn, onto an inert material, such as Al2O3
(IPCC, 2005; Johansson et al., 2006c). A signicant part of the
experimental testing has been performed in laboratory batch
reactors, where particles are subjected to alternating cycles of
oxidizing and reducing conditions (Adnez et al., 2004; Dueso
et al., 2012; Ishida et al., 1996; Mattisson et al., 2008). The CLC
technology has been successfully demonstrated for methane or
natural gas combustion in different prototypes of interconnected
uidized beds, ranging from 0.3 to 50 kWth with over several
hundred combined hours of operation (Abad et al., 2006; Adnez
et al., 2009, 2006; Johansson et al., 2006a, 2006b; Linderholm
et al., 2010, 2008; Rydn et al., 2008a, 2006; Ryu et al., 2003;
Sedor et al., 2008; Son and Kim, 2006; Zafar and Mattisson, 2005;
Zafar et al., 2006). Oxygen carriers for CLC and CLR have also been
tested in xed-bed reactors and thermogravimetric analyzers
(TGA), for measuring the intrinsic kinetics and diffusion limitations (Garcia-Labiano et al., 2006; Iliuta et al., 2010; Ishida et al.,
1996; Jin, 2002; Jin et al., 1999; Noorman et al., 2011a, 2011b,
2007; Ryu et al., 2003). In addition, CLR studies have been
performed in continuous circulating-uidized bed reactor systems
(De Diego et al., 2009; Rydn et al., 2006) and xed-bed units
(Jin 2002). Potential drawbacks of implementing chemical-looping
in xed-bed reactors include low oxygen carrier utilization, high
pressure drop, and large temperature uctuations, as well as the
inability to operate using solid fuels (Zhou et al., submitted for
publication). With repeated cycles of oxidation and reduction, the
reactivity of oxygen carriers have been shown to change, as a
result of pore opening, surface restructuring, and spinel formation
(Jerndal et al., 2009; Linderholm et al., 2009; Shulman et al., 2009;
De Diego et al., 2009). It is of scientic importance to understand
how the gradual chemical and/or physical changes in the oxygen
carrier from continuous usage affect the reactivity of the oxygen
carriers and the overall performance of chemical-looping systems,
which will be the subject of interest of the second part of
this work.
Different mathematical models have been developed to simulate
CLC systems. Ishida et al. (1996) applied the unreacted shrinking core
model to interpret the experimental results on kinetics of the NiO
mixed with yttria-stabilized zirconia (YSZ) particle. Comparison of
model prediction with experimental results indicated that the
reduction reaction was controlled by chemical reaction, while oxidation was controlled by both chemical reaction and ash-layer diffusion
(Ishida et al., 1996; Mattisson et al., 2006). Iliuta et al. (2010) used an
isothermal dynamic plug-ow model to estimate kinetic parameters

Fig. 1. Schematic overview of chemical-looping combustion and reforming.

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

that best t their xed-bed micro-reactor. The gassolid reaction


parameters were then incorporated into a detailed hydrodynamic
model to successfully model their uidized-bed reactor. Sahir et al.
(2011) used the law of additive reaction times to approximate the
gassolid reactions for analyzing the copper oxidation process. Ortiz
et al. (2012) studied the catalytic kinetics of steam methane reforming and watergas-shift for NiO/-Al2O3 and NiO/-Al2O3, and found
that the activity increased with reduction of the oxygen carrier.
Recently, Zhou et al., 2013 conducted a detailed kinetic analysis for
the relevant oxygen carrier reduction reactions of NiO by CH4 and Ni
catalyzed reactions and validated the proposed kinetic scheme with
different experimental results from literature (Zhou et al., 2013).
Temperature variations inside the particle and reactor have
been considered in numerical models (Garca-Labiano et al., 2005;
Noorman et al., 2011a, 2011b, 2007). Noorman et al. (2011a)
proposed a detailed particle model which incorporated the effects
of reaction kinetics, internal and external mass and heat transfer
to study the reduction and oxidation characteristics of oxygen
carriers. Modeling of the CuO/Cu system showed that internal
mass transfer by was hindered by Knudsen diffusion for a range of
pore sizes and instances where relatively large temperature
differences can exist between the particle and the bulk gas
(Noorman et al., 2011a). A more accurate description of diffusion
processes was obtained by applying MaxwellStefan diffusion;
however Fickian diffusion can be reasonably applied in cases with
simple dilute mixtures. Garca-Labiano et al., 2005 analyzed the
reduction and oxidation characteristics of the major oxygen
carriers based on Cu, Co, Fe, Mn, and Ni using a non-isothermal
changing grain size particle model. The temperature changes
inside the particle during the exothermic oxidation reactions
depended on the particle size, overall reaction time, and the
resistance to heat transfer in the external gas lm, with the
highest temperature variation corresponding to Ni for 1 mm
particles (Garca-Labiano et al., 2005).
Heterogeneous modeling can be applied to CLC processes to
investigate the competing effects of uid transport and chemical
reactions occurring within a xed-bed reactor. Oxygen carriers
with particle size ranging from 0.08 to 2 mm have been used in
the literature (Lyngfelt et al., 2001). Larger particle sizes are
advantageous in minimizing pressure drop, but the effects of
intraphase mass and heat transfer limitations can be signicant.
Noorman et al. (2011a) found that particles larger than 0.3 mm are
severely diffusion-limited, which if neglected, models can lead to
unreliable kinetic parameters. The role of heterogeneous modeling
can identify where the improvements to oxygen carrier performance are feasible in the regime of kinetic control or mass transfer
control, thereby aiding oxygen carrier and reactor design. This
work proposes a heterogeneous model that takes into account the
effects of reaction kinetics in association with mass and heat
transport in the particle phase, and convection, dispersion, and
pressure drop in the bulk uid phase. The effects of reaction and
transport limitations on chemical-looping systems operating in
xed-bed reactors are analyzed based on published experimental
results from three bench-scale xed-bed reactors.

2. Model description
Flow in xed-bed reactors can be modeled by the motion of
gaseous bulk uid through the gaps between and within the oxygen
carriers that constitute the solid phase (Avci and Ilsen, 2011). Heat and
mass transfer resistances can occur within the solid phase (intraparticle resistance) or between the solid and the bulk uid (interfacial or
interparticle resistance). Intraparticle resistance is related to the pore
size and can be controlled by the dimensions of the oxygen carrier.
Interfacial resistance is proportional to the thickness of the boundary

235

layer at the gassolid interface. The boundary layer can exhibit


pronounced temperature gradients in cases of high heats of reaction
that drive rapid changes in surface temperatures and concentrations.
Though, this boundary layer effect can be damped by introducing
higher gas ow rates and smaller packing sizes to induce mixing.
According to the classication widely accepted in the chemical
engineering society, the continuum models can be categorized into:
pseudo-homogeneous and heterogeneous models (Bischoff and
Levenspiel, 1962). In pseudo-homogeneous models, the catalyst surface is assumed to be completely exposed to the bulk uid conditions,
such that any uid-to-particle heat and mass transfer resistances can
be neglected. In contrast, heterogeneous models account for transport
resistances of heat and mass from bulk uid phase to the particle
phase and within the particle. In this section, one- and twodimensional pseudo-homogeneous and heterogeneous models are
presented. The models are general and can be applied to any chemical
looping process for any oxygen carrier. In the simulations, the gas
species considered are CH4, CO, CO2, H2, H2O, and an inert (usually Ar)
and the solid species considered are NiO, Ni, deposited carbon, and an
inert (Al2O3). Sections 2.1 and 2.2 detail the continuum models and
Section 2.3 the momentum balance equation. Sections 2.4 and 2.5
summarize the transport correlations, physical properties, and kinetic
scheme chosen for this work.
2.1. Homogeneous model design equations
2.1.1. One-dimensional
The one-dimensional pseudo-homogeneous model describes
the axial proles of the bulk uid concentration and temperature.
The dynamic axially dispersed plug ow reactor model is


C F

C
b i i b Dax;i i b Ri ;
1
z
t V
z

b C p;f C T



T
T

T
C p;f F T
b
b  H i Ri
ax
t
V
z
z
U
 4 T  T w ;
dt

with Danckwerts boundary conditions, expressing continuity of


uxes at steady state:

b Dax;i

C i
j
F i jz 0  yi;f eed F in =Ac ;
z z 0

T
z

b ax jz 0 Tjz 0  T in C p;f F T =Ac ;

C i
T
j
j
0:
z z L z z L

2.1.2. Two-dimensional
Advancing in complexity, the two-dimensional model takes
into account radial diffusion of bulk uid and temperature.
Reaction rates along the reactor diameter can be affected by
effective thermal diffusivity and conductive heat transfer at the
wall, which are determined by ow characteristics. The twodimensional pseudo-homogeneous model is




C F

C

C
rDrad;i i b Ri ;
b i i b Dax;i i b
6
z
t V
z
r r
r

b C p;f C T





T
T

T

C
C p;f F T
b
b
r rad i
ax
t
V
z
z
r r
r
b  H i Ri ;

with boundary conditions at r R and r 0:


C i
T
C
j
j
i jr R 0 ;
r r 0 r r 0
r

236

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

T
r

b rad jr R  hw Tjr R  T w ;

phase are

and at z L and z0 are consistent with Eqs. (3)(5).





C i F i

C

C
Dax;i i b
rDrad;i i kc;i av C c;i jRp  C i ;

b
z
t V
z
r r
r
17

2.2. Heterogeneous model design equations


The presented pseudo-homogenous model is applicable supposing there is negligible difference between solid and uid
conditions and mild radial temperature and concentration
proles. In the event that interparticle differences are important,
a heterogeneous model should be used in place of the pseudohomogenous model. Furthermore, intraparticle resistance is considered because limitations to diffusive transport inside the pores
can signicantly limit the reaction rates inside large particles. The
most important assumptions made are (i) spherical or cylindrical
oxygen carrier, (ii) constant volume of the particle, (iii) macroscopically uniform particle structure that is not affected by reaction, and (iv) uniform Ni distribution within the particle. Local
variations in structure or composition are not taken into the
account due to the absence of experimental data. The gas concentrations and temperature inside the pores of the particle are
functions of oxygen carrier radial direction, rc, and/or axial direction, zc. It is also assumed that the gas at any position inside the
particle is at the same temperature as the solid, and that the
thermal conductivity of the gas phase is negligible compared to
that of the solid. Below, the conservation equations corresponding
to a spherical oxygen carrier are discussed. For cylindrical oxygen
carriers, the particle phase equations are summarized in Appendix.
Note that the transport of uid phase is unspecic to particle
geometry.
2.2.1. One-dimensional
The one-dimensional heterogeneous model for the uid phase
is


C F

C
b i i b Dax;i i kc;i av C c;i jRp  C i ;
10
z
t V
z

b C p;f C T

T
T

T
C p;f F T
b
hf av T c jRp  T;
ax
t
V
z
z

and for the particle phase is




C
C
1
c c;i 2
De;i r c 2 c;i s Ri ;
t
r c
r c r c

11

12



T c
s
T c
rc 2
2
1  c s C p;s c C p;c C T;c
s  H i Ri :
t
r c
r c r
13
The boundary conditions for the particle phase are
C c;i
T c
j

j
0;
r c rc 0 r c rc 0

De;i
 s

14

C s;i
jRp kc;i C c;i jRp C i ;
r


T s
j hf T s jRp  T;
r Rp

15

16

and for the uid phase the boundary conditions are equivalent to
the one-dimensional homogeneous model (Eqs. (3)(5)).
2.2.2. Two-dimensional
The two-dimensional heterogeneous model includes radial
dispersion in the uid phase. The design equations for the uid

b C p;f C T





T
T

T

C
C p;f F T
b
b
r rad i
ax
t
V
z
z
r r
r
hf av T c jRp  T;

and for the particle phase are




C
C
1
c c;i 2
De;i r c 2 c;i s Ri ;
t
r c
r c r c
1  c s C p;s c C p;c C T;c

18

19



T c
s
T c
rc 2
2
s  H i Ri ;
t
r c
r c r
20

with Eqs. (19) and (20) as boundary conditions for the solid phase
and Eqs. (3)(5) for the bulk phase.
2.3. Momentum balance
The momentum balance assumes pseudo-steady state and is
dominated by the friction factor. The pressure drop correlation
using the Ergun friction factor for gas ow through a packed
tubular reactor is written as (Wilkes, 1999)


 2 
u0 150
dP
1  b

1:75
21
3
dz
Rep
Dp
b
where P is the total pressure in the bed and Rep is the particle
Reynolds number, dened as
Rep

u 0 dp
:
1  b

22

The boundary condition at the bed exit is applied to maintain


the efuent stream leaving the reactor at atmospheric conditions.
Pjz L P out

23

2.4. Meat and mass transfer correlations


The effective axial coefcient can be estimated by (Bischoff and
Levenspiel, 1962)
1
b
0:45

Pea Rep Sc 1 7:3=Rep Sc

24

and the effective radial dispersion coefcient can be estimated by


(Foumeny et al., 1992)
1
0:38
0:08
:

Per Rei Sc0:80 1 11:1=Rei Sc

25

In general, for gases with Rep 4 1, Per  10 and Pea  1, and for
gases with Rep o1, Per  Pea which approach 0.7 (Davis and Davis,
2003). Dispersion is usually more dominant in the radial direction,
and consequently very small radial variations of concentration can
be expected in the reactor. However a high dependency of reaction
rates on temperature can lead to pronounced differences between
the one-dimensional and two-dimensional models.
The effective radial and axial thermal conductivity can be
estimated by Eqs. (26) and (27), respectively, by Yagi and Wakao
(1959) and Yagi et al. (1960) for metal spheres. The static
0
contribution, e , is assumed to be 0.01 W/m K.

rad
13 0:11 Pr Rep ;
m

26

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

ax 0e

0:7Pr Rep :
m m

27

The wall heat transfer coefcient is calculated from (Wako


et al., 1979)
Nu 2 1:1Pr 1=3 Rep 0:6 :

28

The bed heat transfer coefcient is calculated by (Peters et al.,


1988)
hbed

4er
:
D=2

29

The resistances to the radial heat transfer in the bed and at the
wall can be combined to dene the overall heat transfer coefcient
to be used in the one-dimensional models:
1
1
1

:
U hbed hw

30

The mass and heat transfer coefcients are estimated as


(Sattereld, 1981):


G
;
31
kc;i 0:357Re  0359 Sc  2=3



0:357
hf 1:37
C p;f G Re  0359 Sc  2=3 :

32

The bed porosity for a spherical catalyst and surface area


per unit volume are calculated by Eq. (33) (Pushnov, 2006) and
Eq. (34) (Fogler, 2005):

1:0
Dt =Dp 2

0:375;

av 61  b =Dp :

33
34

The pore diameter of a cylindrical pore can be calculated by


(Sattereld, 1981)
rp

S g s

Typical NiO oxygen carriers used in CLC processes have void


fraction of 0.30.5, particle size of 0.081.0 mm, and solid density
of 17004000 kg/m3 (Lyngfelt et al., 2001). The pore tortuosity, if
no data is available, can be assumed 3.0 for spherical catalysts
(Marrero and Mason, 1972). The gas phase diffusivities in the
mixture depend on the combined effect of molecular diffusion and
Knudsen diffusion for the relevant range of pore sizes of microporous or mesoporous oxygen carrier particles, typically being
20500 . Binary gas phase diffusivities for a wide range of
temperatures and pressures applicable are estimated through
published correlations detailed in Tables 1 and 2. For all other
component pairs, the semi-empirical Fuller equation can be used
(Poling et al., 2000).
Gas phase diffusion Dm;i can be approximated using (Perry
and Green, 1997)
Dm;i

1  yi
:
j a i yj =Dij

Table 1
Binary gas phase diffusivities for component pairs (Marrero and Mason, 1972).
Pair

Eq.

H2CO
H2CO2
COCO2
CO2H2O
H2CH4
ArH2
ArCH4
ArCO
ArCO2
N2H2O

15.39E  3
3.14E  5
0.577E  5
9.24E 5
3.13E  5
23.5E  3
0.784E  5
0.904E  5
1.74E  5
0.187E  5

1.548
1.75
1.803
1.5
1.765
1.519
1.785
1.752
1.646
2.072

0.316E8

0.488e8

 2.80
11.7
0
307.9
0
39.8
0
0
89.1
0

1067

1
2
2
2
2
1
2
2
2
2

are for these expressions: (1) Dij


Note
21. Parameters
 D12 D
 Bbelow



exp  DT  TE2 ; (2) Dij ATP exp  DT .
Note: T is in K, P is in atm., and Dij is in cm2/s.

36

The effective diffusivity De;i calculates the true diffusion path


from the molecular diffusivity and Knudsen diffusivity DK;i , as
follows:


1
1
1
:
37

De;i b Dm;i DK;i


The gas phase heat capacity (Cp,f) is taken as the molar average
of the gas component heat capacity (Cp,i), which is estimated using

AT B
P

    2
ln CT

Table 2
Binary gas diffusivities for component pairs (Poling et al., 2000; Sattereld, 1981).
Pair

H2N2
CON2
H2H2O
CH4H2O
H2Ar
CH4CO2

6.007E  3
0
0
0
9.625E  3
0

 0.99311
0.322
0.927
0.361
 2.287804
0.153

Note D12 D21. Parameters below are for the expressionDij AT B=P.
Note: T is in K, P is in atm, and Dij is in cm2/s.

Table 3
Gas phase heat capacity constants for relevant species, corresponding to the
equation C p;i A BC=T sinh C=T2 DE=T sinh E=T2 (Aspen, 2007).

35

2.5. Physical properties

237

CH4
H2
H2O
CO
CO2
Ar
N2

33,298
27,617
33,363
29,108
29,370
20,786
29,105

79,933
9560
26,790
8773
34,540

8614.9

2086.9
2466
2610.5
3085.1
1428

1701.6

41,602
3760
8896
8455.3
26,400

103.47

991.96
567.6
1169
1538.2
588

909.79

Note: T is in K and Cp,i is in J/kmol K.

Table 4
Gas phase viscosity constants for selected species, corresponding to the equation
mi AT B =1 C=T D=T 2 (Aspen, 2007).

CH4
H2
H2O
CO
CO2
Ar
N2

5.255E  07
1.797E 07
1.710E  08
1.113E  06
2.148E  06
9.212E  07
6.5592E  07

0.59006
0.685
1.1146
0.5338
0.46
0.60529
0.6081

105.67
 0.59
0
94.7
290
83.24
54.714

0
140
0
0
0
0
0

Note: T is in K and is in N s/m2.

the DIPPR equation from Aspen Plus software and shown in


Table 3 (Aspen, 2007).
The gas phase viscosity of species can be estimated from the
DIPPR correlation used in Aspen Plus, shown in Table 4. The
viscosity of the gas mixture is estimated by the Wilke method

238

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

with Herning and Zipperer approximation (Poling et al., 2000):

NoComp

i1

yi i
p:
NoComp
yj M j =M i
j1

38

The gas phase thermal conductivity of species is estimated with


the DIPPR equation used in Aspen Plus (Aspen, 2007), shown in
Table 5. The thermal conductivity of the mixture is calculated by
the Wassiljew equation (Poling et al., 2000),

NoComp

yi i

i1

NoComp
yj Aij
j1

210

T c M3

!1=6
42

P 4c

The thermal conductivity of the oxygen carrier is calculated


from the volume fraction of the metal compounds in the particle
(Eq. (43)), and thus varied with the conversion (Garca-Labiano
et al., 2005). Table 6 presents the physical properties of solids
commonly used as oxygen carriers.

1 f v;i
:
s;i

43

39
2.6. Kinetic scheme

The binary interaction parameter (Aij) can be found based on


Mason and Sexena's formulation (Eqs. (40)(42)), where the ratio
of the translational thermal tri =trj conductivities needs to be
calculated from the reduced temperatures (Tr) and reduced,
inverse thermal conductivity () (Poling et al., 2000).
Aij

1 tri =trj 1=2 M i =Mj 1=4 2

40

81 M i =M j 1=2

tri j exp0:0464T ri exp  0:2412T ri 

trj i exp0:0464T rj exp  0:2412T rj 

41

Table 5
Gas phase thermal conductivity constants for species, corresponding to the
equation AT B =1 C=T D=T2 (Aspen, 2007).

CH4
H2
H2O
CO
CO2
Ar
N2

8.398E  06
2.653E  03
6.204E  06
5.988E  04
3.69
6.33E 04
3.31E  04

1.4268
0.7452
1.3973
0.6863
 0.3838
0.6221
0.7722

 49.654
12
0
57.13
964
70
16.323

0
0
0
501.92
1.86E 06
0
373.72

The main reduction reactions in CLC of NiO with CH4 fuel are
shown in Table 7 (Zhou et al., 2013). The reducing gases that are
active in the oxygen carrier reduction reactions (R1R4) comprise
CH4, H2, and CO. In the work by Zhou et al. (2013) and Iliuta et al.
(2010), non-catalytic gassolid reactions have been analyzed using
the shrinking core model and a modied volumetric model.
Because the modied volumetric model has been found to
improve the model predictions, the kinetics based on the modied
volumetric model is used in this work. The kinetic expressions for
the four NiO reduction reactions are shown below, where a0 is the
initial oxygen carrier surface area, X is the oxygen carrier conversion, ks1ks4 are the Arrhenius rate constants, and C denotes the
concentration of the gas species and solid weight fraction of NiO
and Ni.
r 1 a0 1  Xk1 C CH4 C NiO C Ni

44

r 2 a0 1  Xk2 C H2 C NiO

45

r 3 a0 1  Xk3 C CO C NiO

46

r 4 a0 1  Xk4 C CH4 C NiO C Ni

47

The conversion rate for the oxygen carrier is written as


0
dX=dt 2r 1 r 2 r 3 r 4 =C NiO

Note: T is in K and is in W/(m K).

48

Table 6
Physical properties of solids (Garca-Labiano et al., 2005).
Thermal conductivity
(W/m K)
Active material
Co
CoO
Cu
CuO
Cu2O
Fe
FeO
Fe2O3
Fe3O4
MnO
Mn2O3
Mn3O4
Ni
NiO
Inert
Al2O3 ()
MgO
SiO2
TiO2 (rutile)
ZrO2

Heat capacity, Cpa (J/kg K)


Cp0

Solid density, s,i

Cp1

Cp2

Cpb

Melting temperature

(kg/m )

(K)

40
5
350
6

636
751
382
430

 5.13E  01
 1.06E  01
7.46E  03
4.26E  01

5.03E  04
1.03E  04
6.76E  05
 1.53E  04

761
775
460
705

8860
6440
8920
6400

30
5
5
6
3
3
3
75
13

433
623
860
2009
546
498
490
347
790

1.45E  01
2.72E 01
2.96E  03
 1.77
3.12E  01
5.08E  01
4.90E  02
2.71E  01
 2.06E  01

1.66E  06
 6.62E  05
1.57E  05
6.70E 04
 8.80E  05
 1.29E  04
 1.28E  04
 5.52E  05
1.43E  04

613
856
887
847
796
926
358
596
752

6980
5700
5240
5180
5430
4500
4840
8900
6670

1768
2078
1358
NDPT 1397
1516
1809
1650
NDPT 1735
1870
2115
NDPT 1350
1835
1726
2228

7
8
2
4
3

843
1013
953
785
482

5.40E  01
3.47E  01
2.91E  01
2.12E  01
1.74E  01

 1.61E  04
 8.96E  05
 7.82E  05
 5.97E  05
 4.09E  05

1263
1304
1192
955
634

3965
3580
2260
4260
5600

2290
3105
1996
2130
2950

Note: Cpa Cp0 Cp1 T Cp2 T 2 ; Cpb at Tref 1223 K (expect for Cu based OC where Tref 7073 K); NDPT normal decomposition point.

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

239

Table 7
CLC reduction reactions using NiO oxygen carrier selected by Zhou et al. (2013).
Oxygen carrier reduction reactions
Partial CH4 oxidation
H2 oxidation
CO oxidation
Partial CH4 oxidation

CH4 2NiO-2Ni CO2 2H2


H2 NiO-Ni H2O
CO NiO-Ni CO2
CH4 NiO-Ni 2H2 CO

(R1)
(R2)
(R3)
(R4)

Reduction reactions catalyzed by Ni


Steam methane reforming
Water gas shift
Dry reforming
Methane decomposition
Carbon gasication by steam
Carbon gasication by CO2

CH4 H2O23H2 CO
CO H2O2H2 CO2
CH4 CO222CO2H2
CH422H2 C
C H2O2CO H2
C CO222CO

(R5)
(R6)
(R7)
(R8)
(R9)
(R10)

Table 8a
Summary of independent variables in the one-dimensional model for a spherical oxygen carrier.
Variable
C(z,i)
F(z,i)
T(z)
P(z)
Cc(z,rc,i)
Tc(z,rc)

Unit
3

mol/m
mol/s
K
bar
mol/m3
K

Description

Dimension

Gas phase concentration of species i


Gas phase molar ow rate of species i
Gas phase temperature
Gas phase pressure
Concentration of i in the gas inside the catalyst pores at z
Temperature of the catalyst at z

Nz  Ns
Nz  Ns
Nz
Nz
Nz  Nrc  Ns
Nz  Nrc

Table 8b
Summary of independent variables in the two-dimensional model for a spherical oxygen carrier.
Variable

Unit

Description

Dimension

C(z,r,i)
F(z,r,i)
T(z,r)
P(z,r)
Cc(z,r,rc,i)
Tc(z,r,rc)

mol/m3
mol/s
K
bar
mol/m3
K

Gas phase concentration of species i


Gas phase molar ow rate of species i
Gas phase temperature
Gas phase pressure
Concentration of i in the gas inside the catalyst pores at z
Temperature of the catalyst at z

Nz  Nr  Ns
Nz  Nr  Ns
Nz  Nr
Nz
Nz  Nr  Nrc  Ns
Nz  Nr  Nrc

The mass balance for Ni and NiO is (assuming no migration of


the Ni particles)
0
dC NiO =dt C NiO
1  XM NiO
0
X
dC Ni =dt C NiO

49

55

50


 
2
P CO2 P CO
P CO2
k10

= 1 K CO;10 P CO
r 10
K CO;10 K CO2 ;10 P CO K 10
P CO K CO;10 K CO2 ;10

For the Ni-based system, the catalytic gassolid reactions are


steam methane reforming (R5), water gas shift (R6), dry reforming
(R7), methane decomposition (R8), and carbon gasication by
steam (R9) and CO2 (R10). Normally, the rates of NiO reduction
reactions (R1R4) are dominant initially when NiO is in excess,
while the catalytic reactions become increasingly important as Ni
is formed. The rate expressions chosen by Zhou et al. (2013) are
!
P 3H P CO
k5
r 5 2:5 P CH4 P H2 O  2
K5
PH
2

=1 K CO;5 P CO K H2 ;5 P H2 K CH4 ;5 P CH4 K H2 O;5 P H2 O =P H2 2


r6

51



P H P CO2
k6
P CO P H2 O  2
P H2
K6
=1 K CO;5 P CO K H2 ;5 P H2 K CH4 ;5 P CH4 K H2 O;5 P H2 O =P H2 2

r 7 k7 K CH4 ;dr P CH4 P CO2 

r 8 k8 K CH4 ;8 P CH4 

!
P 2CO P 2H2
=1 K CH4 ;7 P CH4
K7

!
!2
3=2
PH
P 2H2
= 1 2 K CH4 ;8 P CH4
K8
K H2 ;8

!2


3=2
P H2
P H2 O P CO
P H2 O
r9


= 1 K CH4 ;9 P CH4
K H2 O;9 P H2
K9
P H2 K H2 O;9 K H2 ;9
k9

56
2.7. Integration strategy
The dynamic model was implemented and solved in the
equation-based process modeling platform gPROMS 3.6.0
(Process Systems Enterprise Limited, 2012). The nite difference
method is used for the discretization of partial differential equations along the xed-bed axial (Nz nodes) and radial (Nr nodes)
direction, and along the solid radial and/or axial directions (Nrc
and Nzc nodes). The number of species is denoted as Ns. A
summary of independent variables can be found in Tables 8a
and 8b. In the proceeding sections, the results are presented and
discussed.

52
3. Results and discussion
53

54

The developed pseudo-homogeneous and heterogeneous models are applied to published chemical-looping xed-bed experiments. In Section 3.1, the model predictions are shown against the
experimental data, where the differences in estimated kinetic
parameters will be discussed based on the importance of various

240

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

transport effects and the assumptions associated with the model.


The effect of dispersion and radial temperature variations is
analyzed for each discussed experiment in Sections 3.2 and 3.3.
A detailed analysis for the effect of internal and external diffusion
resistance on individual reaction rates is conducted in Section 3.4
using the WesizPrater and Mears numbers. Lastly, in Section 3.5
the oxygen carrier properties are treated as variables in an
optimization scheme to select an optimal oxygen carrier for CLR
that maximizes CH4 conversion and H2 selectivity.
The proposed models are applied to predict experimental
results from Iliuta et al. (2010), Jin (2002), and Rydn et al.
(2008b), whose xed-bed reactor designs and operating conditions are summarized in Table 9. These experiments utilize
methane as the fuel and a Ni-based oxygen carrier supported over
alumina, which is within the scope of the kinetic scheme. In
general, the xed-bed reactors are housed in an electrical furnace
and operated at atmospheric conditions, although Jin (2002) also
conducted high pressure experiments for CoNi/YSZ oxygen carriers. Iliuta et al. (2010) and Rydn et al. (2008b) operated in
chemical-looping combustion conditions at 900 1C, while Jin
(2002) operated in chemical-looping reforming conditions at
600 1C.
The oxygen carrier properties varied in size, preparation,
support material, surface area, and pore size. Surface area measurements for the oxygen carriers used by Jin (2002) and Rydn
et al. (2008b) were unavailable so they were estimated from Fan
(2011) and Mendiara et al. (2011) accordingly. In the experiment
by Rydn et al. (2008b), b was estimated from Eq. (33) to be
approximately 0.4 and b was assumed to be 1000 kg/m3. The
porosity of all of the studied oxygen carriers (c) was assumed to
be 0.37. To match experimental data, dynamic parameter estimation is performed to t the frequency factor of the kinetic rate
constant of all the reactions. The model assumes a constant wall
temperature (of the reported furnace setting), atmospheric pressure at the reactor outlet, uniform distribution of oxygen carrier
particles inside the bed, and homogeneity of the oxygen carrier
properties (including particle size, pore size, and surface area).
Initially, inert gas is present inside the reactor, and then following
t0 , the reactant gas ows into the reactor to commence the
chemical looping reduction reactions. The simulation is run until
complete reduction of the oxygen carrier is achieved. By applying
the models to various experiments, a wide range of space times

and particle sizes was studied and various contributing effects of


transport limitations and dispersion were analyzed. The estimated
kinetic parameters are presented in Tables 1012.
3.1. Effect of particle size on diffusion limitation
The importance of mass and transport resistance inside the
particle can be evaluated based on the particle size. Since nearly all
of the active sites are located within the pore network, the
Table 10
Estimated kinetic rate constants for modeling CLC experimental data reported by
Iliuta et al. (2010) at 900 1C.
Frequency factors

1D homogeneous model

1D heterogeneous model

k1
k2
k3
k4
k5
k6
k7
k8a
k9
k10

1.63E  05
2.29E  05
1.42E  05
2.29E  05
3.54E  03
1.64E  05
5.41E  05
5.10E  05
5.40E-07
1.35E-04

8.16E  06
2.26E  05
1.32E  05
2.13E  05
5.44E  03
1.68E  05
9.37E  05
6.07E  05
5.40E-07
1.07E-04

kCH4;8 is estimated in lieu of k8.

Table 11
Estimated kinetic rate constants for modeling CLR experimental data reported by
Jin (2002) at 600 1C.
Frequency factors

2D homogeneous model

2D heterogeneous model

k1
k2
k3
k4
k5
k6
k7
k8a
k9
k10

1.46E  05
4.38E  07
2.05E  07
1.61E  08
4.70E 07
8.56E  04
7.35E 09
0.00E 00
1.06E  08
2.38E  04

5.84E  06
1.26E  07
6.41E  07
8.50E  08
1.82E  06
3.10E  02
7.34E  06
1.62E  03
8.71E  04
1.96E  04

kCH4;8 is estimated in lieu of k8.

Table 9
Oxygen carrier physical properties and experimental conditions of chemical-looping xed-bed units studied in this work.

OC properties

Lbed (mm)
Dreactor (mm)
Q (ml/min)
U (W/(m2 K))
(kg/m3)
(1e5  Pa s)
Cpf (J/(kg K))
m, ax, rad [1E2  W/(m K)]
Dax, Drad, De (1E5  m2/s)
Re
Sc
Pea, Per
Pr [x1E2]
Sh
a
b

Estimated from Table 3.17 from Fan (2011).


Estimated from Mendiara et al. (2011).

Iliuta et al. (2010)

Jin (2002)

Rydn et al. (2008b)

NiO/Al2O3 (15 Ni wt%)


Dp 140E  3 mm
a0 102 m2/g
eb 0.37
b 1040 kg/m3
7.65
4
100 (10% CH4 in Ar)
480880
0.350.46
5.86.2
2027
4.99.1; 1.01.1; 63118
110; 8.064; 0.0280.13
0.420.52
0.171.8
0.635.4; 0.382.4
1.62.7
0.350.76

NiO/NiAl2O4 (60 Ni wt%)


Dp 4.0 mm; Lp 1.5 mm
a0 10 m2/ga
eb 0.487
b 995 kg/m3
250
16
900 (H2O/CH4 2)
30110
0.20.6
2.95.1
2050
417; 1.51.8; 52200
3263; 1550; 0.582.4
610
0.22.3
35; 812
0.740.27
1.63.8

NiO/MgAl2O4 (60 Ni wt%)


Dp 120E-3 mm
a0 2 m2/gb
b 1000 kg/m3
5.659
15
60 (25% CH4 in Ar)
240800
0.30.4
5.36.2
2035
516; 1.0; 63210
1.010; 320; 0.423.0
0.0180.021
0.23.2
0.020.4; 0.030.2
0.942.65
0.411.0

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

diffusion of molecules plays a crucial role in the observed rate of


reaction. Molecular diffusion cannot capture the internal transport
alone because the mean-free path of the gases often exceeds the
pore length of the typical solid catalyst. Noorman et al. (2011a)
found Knudsen diffusion to be very relevant for chemical-looping
Table 12
Estimated kinetic rate constants for modeling CLC experimental data reported by
Rydn et al. (2008b) at 900 1C.
Rate constants

1D homogeneous model

2D homogeneous model

k1
k2
k3
k4
k5
k6
k7
k8a
k9
k10

3.65E  04
2.93E  05
2.20E  05
3.49E  06
6.30E  04
2.41E 01
1.59E  04
1.14E  02
5.56E  06
1.92E  02

3.39E  04
2.89E  05
3.21E  05
3.60E  06
5.58E  04
2.41E 01
1.59E  04
1.37E  02
5.65E  06
2.38E  02

kCH4;8 is estimated in lieu of k8.

Fig. 2. 1D model predictions of the experimental data reported by Iliuta et al. (2010).

oxygen carriers with micro- and meso-porosity. As a guideline,


diffusional limitations can be neglected for particles below 0.2 mm
in diameter (Garca-Labiano et al., 2005). For larger particles
(Dp 40.2 mm), there is an inuence of both chemical reaction
kinetics and internal diffusion that can be captured with detailed
heterogeneous modeling. Two case studies are presented to
analyze the effect of particle size on overall transport resistance
and determination of kinetic parameters.
In the xed-bed experiment conducted by Iliuta et al. (2010),
the particle size was approximately 140 m and thus, intraparticle
and interfacial effects were neglected in their modeling approach.
Here, the one-dimensional homogeneous and heterogeneous
models are used to predict the experimental exit concentrations.
There is no signicant improvement from using a twodimensional model to characterize this reactor because radial
dispersion effects are minimal (discussed in Section 3.2). It can
be seen in Fig. 2 that both models are capable of predicting the
experimental data and hardly any differences can be observed
between the two models. Thus, the quality of t, expressed as the
sum of squared errors, is relatively unchanged. The current
modeling predictions are superior to the original t published by
Iliuta et al. (2010) for the 900 1C xed-bed experiments because of
the inclusion of energy balance, momentum balance, axial dispersion, and molar expansion in the design equations. The error in the
H2O prediction and experimental result, which is evident in the
original t and in Fig. 2, can be explained by the calibration
difculty of H2O on the mass spectrometer and the condensation
of H2O prior to the mass spectrometer (Iliuta et al., 2010).
The effect of diffusional resistance on individual reaction rates
is analyzed using the heterogeneous model. By virtue of the
effectiveness factor, a quantitative measure can be obtained for
the importance of diffusion to reaction limitations. Reactions that
are limited by kinetics exhibit an internal effectiveness factor
approaching 1 and diffusion-limited cases observe 1. In nonisothermal conditions, values of can be greater than 1 because
the external surface temperature is less than the internal temperature of the solid where an exothermic reaction is taking place.
Fig. 3 shows the variation of the effectiveness factor with time in
the reduction cycle. As expected from the small particle size, the
reaction rates are essentially unaffected by diffusional limitations.
R2 is the least diffusion-limited reaction among the NiO reduction
reactions because of the high effective diffusivity of H2. The
effectiveness factors of R1 and R4 decrease to a minimum at
t4 seconds, while that of R2 and R3 see a maximum. This is
attributed to the sensitivity of the reaction rates to the temperature drop inside the reactor (and particle), since R2 and R3 are
exothermic and R1 and R4 are endothermic. The Ni-catalyzed

2.25

5
R1 and R4
R2
R3

R5
R6 (reverse)
R7
R8
R9
R10

Effectiveness factor

Effectiveness factor

1.75

1.5

1.25

10

241

Time (seconds)

10

10

Time (seconds)

Fig. 3. Effectiveness factors for R1R10 from the 1D heterogeneous model for the xed-bed reactor reported by Iliuta et al. (2010).

242

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

Gas concentration (dry basis)

reactions are kinetically-controlled because the calculated effectiveness factors are one or higher. The negligible effect of diffusional limitations inside the small particle supports the close
agreement between the homogeneous and heterogeneous simulations in terms of exit concentration proles (Fig. 2) and estimated
kinetic parameters (Table 10).
In comparison, resistance to mass and heat transfer is important for large particles (Dp 4 0.2 mm), where the local concentration and temperature inside the particle deviate from the surface.
The model needs to distinguish between the uid and solid phase
to adequately account for the internal and external diffusion into
the particle. The CLR xed-bed experiment conducted by Jin
(2002), who prepared cylindrical-shaped oxygen carriers of
4.0 mm diameter and 1.5 mm length, is modeled using the
developed heterogeneous and homogeneous models. Radial variations within the structure of the xed-bed are accounted for in

Times (minutes)
Fig. 4. 2D model predictions of the experimental data reported by Jin (2002).

the two-dimensional models, described in Sections 2.1.2 and


2.2.1. The bed of oxygen carriers is assumed to be uniform
cylindrical pellets situated vertically and stacked in rotating
layers to avoid gas by-pass. While the homogeneous model does
not explicitly account the presence of the oxygen carrier, the
heterogeneous particle model considers lm resistance at the
surface and internal diffusion along the axial and radial directions of the particle.
Fig. 4 shows the necessity of accounting for intraparticle and
interfacial limitations to achieve accurate simulations and predictions.
Compared to the homogeneous model, the heterogeneous model
achieves over 30% improvement in quality of t in terms of sum of
squared errors, and noticeable improvements are seen in the CH4, CO2,
and H2 proles. Large model-experimental deviations observed for the
rst 2 min can be explained by possible dilution of the exiting gas
between the reactor exit and gas chromatographermass spectrometer (GCMS), which is not included in the xed-bed reactor model.
The homogeneous model over-predicts the oxygen carrier conversion,
which leads to higher-than-experimental H2 and CO2 concentrations.
The variation of effectiveness factor for all reactions with reduction
time is shown in Fig. 5. In the rst 5 min, the rates are low due to
methane activation, so transport within the particle is not the limiting
factor and high effectiveness factors are observed. Beyond t 5 min,
reaction rates are developed and conversion of the oxygen carrier
begins. Effectiveness factors of the NiO reduction reactions remain
relatively at one, indicating a high driving force for NiO reduction.
However, diffusional limitations are observed for the catalytic reactions for which effectiveness factors reach values less than one. This is
related to the conversion of the solid NiO material to metallic Ni,
which initiates at the external surface of the particle, and progresses
toward the interior of the particle. Catalytic reaction rates are thus
affected by the Ni concentration and tend to be highest on the surface,
decreasing progressively along the interior. Local gas concentrations
inside the particle are non-uniform and tend to approach equilibrium
values moving away from the surface. When the reactions are close to
equilibrium, the rates are sensitive to changes in local concentration
gradients, and reverse reactions can be become favored. A likely cause
for the rate reversal of R6 and R7 is attributed to the accumulation of
combustion and reforming products (H2, CO, CO2) due to R1R4 inside
the particle that is re-directing the position of equilibrium reactions.
Because of the strong particle effects, estimated rate constants vary
signicantly depending on the model, as shown in Table 11. The
homogeneous model generally estimates lower kinetic rate constants
because the model compensates for internal transport resistance by
lowering the rates. This demonstrates that diffusion resistance cannot

R1
R2
R3
R4

2.5

Effectiveness factor

2.5

Effectiveness factor

R5
R6
R7
R8
R9
R10

1.5

0.5

1.5

0.5

0
0

10

Time (minutes)

15

20

10

15

20

Time (minutes)

Fig. 5. Effectiveness factor for R1R10 from the heterogeneous model for the xed-bed reactor reported by Jin (2002).

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

be neglected and a heterogeneous model is more realistic at predicting the experimental results.
3.2. Effect of dispersion
The presented models are developed to describe cases of nonideal plug ow, whereby components are transported by bulk ow
and convective diffusion (or dispersion). The CLC experiment
performed by Rydn et al. (2008b) is representative of a system
where there is signicant contribution to transport by dispersion.
The low Pclet number and even lower Reynolds number, shown
in Table 9, point towards a mixed reactor. The importance of axial
dispersion can be analyzed by noticeable differences in concentration proles from the dispersive and non-dispersive models
(Fig. 6). Dispersion serves to broaden concentration and temperature gradients and is important to characterize in the direction
of bulk ow. In addition to its strongly dispersive ow,
particle effects are negligible, owing to the small particle size
(Dp 120 m) and large pore size. As shown in Fig. 6, the homogeneous model has been able to adequately predict the experimental data. The heterogeneous model predictions are also
satisfactory, but the model does not bring added benet in terms

of quality of t and thus need not be applied for estimation of


kinetics.
The radial mass dispersion is very fast, so although the reactor
diameter is comparably large, changes to the reaction rate due to
temperature uctuations across the radial direction are quickly
recovered. Therefore, the two-dimensional model predictions do
not deviate signicantly from the one-dimensional model. The
radial dispersion on reaction rates can be evaluated based on the
estimated kinetic rates. As shown in Table 12, the similarities
between the two models further indicate that radial concentration
gradients are not signicant in this system. One note to mention is
that the estimated value for k6 is large, when compared to
literature values. This may be caused by inaccuracy of the H2O
calculation by Rydn et al. (2008b) because only complete oxidation of CH4 was considered signicant in the beginning of the
reduction cycle. However, this is not the case because NiO
reduction proceeds through partial oxidation of CH4 into CO and
H2 and inaccuracies in H2O calculations largely impact the estimated kinetics of R6.
3.3. Radial temperature variations
Detailed modeling of the temperature and conversion patterns
in the xed-bed necessitates inclusion of the resistance to heat
and mass transfer in the radial direction of the uid phase. Radial
temperature gradients that arise from inadequate heat transfer can
effectively alter reaction rates. In the experiment by Rydn et al.
(2008b), temperature variations of around 40 1C can be expected,
in which the largest temperature decrease is exhibited at r 0. The
solid phase temperature is equivalent to the bulk uid, and from
Fig. 7, does not change inside the particle. Interfacial heat and
mass exchange is much higher compared to the reaction rate
because particle effects are not important in this system.
In the other previously discussed CL experiments by Jin (2002)
and Iliuta et al. (2010), the presence of dispersion in axial and
radial directions is less signicant. However, it is interesting to
explore if radial gradients occur. Fig. 8 shows the uid and solid
temperature prole prediction for the experiment conducted by
Iliuta et al. (2010). The temperature drop inside the reactor reaches
up to 45 1C, while the radial temperature variation is at most 30 1C.
The reaction rates are not noticeably affected by the thermal
gradients because of the small temperature uctuation and small
reactor diameter. Therefore, a one-dimensional model is suited to
model this experiment. Fig. 8 also shows that the temperature
inside the particle does not vary along the radius of the oxygen
carrier, due to the high thermal conductivity of the solid. The
difference between the temperature in the solid phase and the
bulk uid temperature is 3 1C, which points to a negligible
resistance for external heat transfer.

900

Catalyst Temperature (C)

Fluid Temperature (C)

Fig. 6. Homogeneous model predictions of the experimental data reported by


Rydn et al. (2008b).

890
880
870
860
6
4
Time (min)

2
0 0

243

8
6
4
Bed Radial Direction (mm)

870
869
868
867

0.02

0.04

0.06

Particle radius, rc (mm)

Fig. 7. Radial uid and solid temperature prole predicted by the 2D heterogeneous model for the experiment conducted by Rydn et al. (2008b).

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

900

Catalyst Temperature (C)

Fluid Temperature (C)

244

890
880
870
860
850
30
20

Time (s)

10
0

868
866
864

1.5

0.5

870

0.02

0.04

0.06

Particle radius, rc (mm)

Bed Radial Direction (mm)

Catalyst Temperature (C)

Fluid Temperature (C)

Fig. 8. Radial uid and solid temperature prole predicted by the 2D heterogeneous model for the experiment conducted by Iliuta et al. (2010).

600

580

560

540
538
536
534
532
530

0.5

1.5

Particle radius, rc (mm)

540
30
20

Time (min)

10
0

Bed Radial Direction (mm)

Fig. 9. Radial uid and particle temperature prole predicted by the 2D heterogeneous model for the experiment conducted by Jin (2002).

The radial temperature prole for the CLR experiment conducted by Jin (2002) is shown in Fig. 9 and predicts a temperature
drop of 58 1C in the reactor and 20 1C along the radial direction of
the bed due to insufcient heat transfer rate with the wall. The
temperature inside the particle remains constant; however, the
particle temperature can be 15 1C colder than the bulk. This
temperature difference results from the interfacial heat transfer
resistance which causes a lower temperature in the particle. Due
to moderate thermal gradients and strong dependency of the
reaction rates on temperature, modelmodel discrepancy between
the one- and two-dimensional models is observed. Radial mass
dispersion limitations arise in the Jin (2002) experiment because,
even though the reactor diameters chosen by Jin (2002) and Rydn
et al. (2008b) are comparable, the radial Pclet number is much
larger than that observed for Rydn et al. (2008b) and contribute
to a smaller dispersion rate, and thus lower heat transfer.

3.4. Investigation of internal and external mass transport effects


In view of heat and mass transfer limitations, the experiment
by Jin (2002) is well represented using the heterogeneous model.
A lack of signicant intraphase diffusion effects on an nth-order
irreversible reaction in a spherical catalyst pellet can be assessed
by the WeiszPrater criterion,
r obs c a2l
De;i C Si

1
{ ;
n

57

where the observed rate, r obs , and the concentration of the


reactant at the external surface C Si are used (Davis and Davis,
2003). This criterion has to be modied for a rst-order reversible

reaction and cylindrical particle,


r obs c a2l
De;i C Si  C EQ
i

{1;

58

is the concentration of the reactant if the reactor were


where C EQ
i
at equilibrium and the coordinates for C Si are taken to be at the
particle radius and mid-particle length.
The inuence of mass transfer through the lm surrounding a
spherical catalyst particle can be examined using a similar expression. Satisfaction of the Mears criterion can be used to determine if
interphase mass transfer is not signicantly affecting the measured rate,
r obs b al 0:15
;
o
n
kf C i

59

in which case the observed reaction rate deviates less than 5%


from the reaction calculated assuming a homogeneous mixture
(Fogler, 2005). The Mears criterion can be modied for a reversible
rst-order reaction in a likewise manner to the WeiszPrater
criterion,
r obs b al
kf C i  C EQ
i

o 0:15:

60

Satisfaction of the Mears and WeiszPrater criteria are determined for each reactant and corresponding reaction. Specically,
Eqs. (57) and (59) are used for R1R4 and Eqs. (58) and (60) for
R5R10. The equilibrium concentrations are calculated from the
equilibrium constants and the corresponding species concentration at the uid or solid phase. The reaction rates of the forward
reactions are integrated over the particle volume to give the
observed rates.
A common nding by investigators is that if mass interphase
gradients occur, then mass intraparticle gradients would occur as

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

10

0.1

0.01

0.1

Mears number

Mears number

103
104
105

104

106
R1
R2
R3
R4

107
108

245

10

15

R5-CH4
R5-H2O
R6-CO
R6-H2O
R7-CH4

105

20

106

10

R7-CO2
R8-CH4
R9-H2O
R10-CO2

15

20

Time (minutes)

Time (minutes)

Fig. 10. Mears numbers reported for R1R7. Fixed-bed unit operated by Jin (2002) is simulated using the heterogeneous model.

10
1

Weisz-Prater number

Weisz-Prater number

0.1

0.01

103
4

10

105
106

0.1
0.01
103
104
105

10

15

20

Time (minutes)

106

10

15

20

Time (minutes)

Fig. 11. WesizPrater numbers reported for R1R10. Fixed-bed unit operated by Jin (2002) is simulated using the heterogeneous model.

well, which both have to be negligible to apply a homogeneous


model (Jakobsen, 2011). The calculated Mears numbers for interphase concentration gradients for the experiment by Jin (2002) are
shown in Fig. 10. The observed rates for the methane conversion
reactions are very slow during the onset of reduction, which
explains the very small Mears numbers at to2 min. For the entire
reduction period, negligible lm transfer resistance is observed for
all the reactions with the exception of CO consumed by R6. Mears
numbers for R1R4 generally decrease over time due to the
decrease in reaction rates from NiO conversion. Because R4 is
the least signicant reaction among R1R4 in terms of rate, Mears
numbers reported for R4 are the smallest. In comparison, the
catalytic reactions observe a gradual increase in Mears numbers
over reduction time, because the reaction rates show no tendency
of diminishing while there are active catalytic surfaces (Ni),
reforming fuel (CH4), and oxygen sources (H2O). The Mears
numbers of R6 reveal that the reaction rate is high at the particle
surface but limited by interphase transport of CO from the bulk,
due to overall low concentrations and small gradients.
In a similar analysis, the importance of internal diffusion
limitations can be evaluated by fulllment of the WesizPrater
criteria, as shown in Fig. 11. A WesizPrater number of greater than
1 indicates that the rate of reaction is much greater than the rate
of diffusion, and that internal diffusion is limiting the overall
reaction scheme. The rates of diffusion of CH4, H2, and H2O

through the pores are generally higher than the rates of CO and
CO2 due to higher effective diffusivities. Therefore, if internal
transport resistance is signicantly controlling a reaction, it is
expected that the slower diffusing reactant will observe higher
WesizPrater numbers. By analysis of Fig. 11, the NiO reduction
reactions satisfy the WesizPrater criteria, and a few catalytic
reactions are limited by internal diffusion. The reaction most
affected by internal transport resistance is R6, with the diffusion
rate of CO to be the most limiting. This nding is expected as lm
resistance is also detected from analysis of the Mears number. R6
is close to equilibrium inside the particle and thus is controlled by
the slow rate of diffusion of CO through the particle. Other
reactions that are limited by internal limitations are the CO2
consuming reactions, R7 and R9, which exhibit slightly greater
than 1 WesizPrater numbers. This is likely due to a slow diffusion
rate CO2 leading to high concentration gradient and non-uniform
reaction rates. As shown, the internal mass transport, as determined by the WesizPrater number, plays a critical role on the
reaction rates, as concentrations deviate much from the bulk.
This analysis reveals the importance of internal and external
mass transfer in chemical-looping reduction reactions. The former
is primarily associated with the oxygen carrier properties, like
particle size and pore size, and the latter to ow characteristics. To
visualize the variation of solid concentration inside a particle at
any given time, the mole fraction of the gas is plotted with respect

246

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

0.21
CO2
CO

0
0...0.02
0.64

H2O

0
0.15
H2

0
0.32

CH4

Fig. 12. Solid concentration variation with particle radius at zp Lp/2. At each
radius, concentration of species is denoted by the color bar where at the surface
represents rp Rp and the center represents rp 0. Fixed-bed unit operated by Jin
(2002) is simulated using the heterogeneous model.

to the radius of the cylindrical particle (Fig. 12). The arc length at
each radial coordinate represents the fraction of each individual
gas species inside the particle (excluding inert gas). For instance, at
positions close to the inlet of the reactor, the surface concentration
is mainly H2O and CH4 because of the direct transfer with the bulk
uid. Approaching rc-0, the conversion of CH4 and H2O and
corresponding production of H2, CO, and CO2 is depicted by the
decreasing arc length of CH4 and H2O and increasing arc length for
H2, CO and CO2. As discussed, intraparticle gradients arising large
particle dimensions and slow diffusion rates lead to non-uniform
reaction rates and departure from optimum performance. In the
next section, the limits due to internal diffusion are explored
through design optimization of the oxygen carrier geometry to
meet chemical-looping standards.

S H2

F H2 ;out
2F CH4 ;in F H2 O;in

Original
Optimal size

Length (mm)

Diameter (mm)

X CH4

S H2

1.5
0.42.0

4
0.8

0.74
0.83

0.52
0.56

61
62

4. Conclusions

As an effective design tool, heterogeneous modeling can be


used to determine an optimal oxygen carrier particle size for a
xed-bed CLR system, in which CH4 conversion and H2 selectivity
are maximized. The dynamic optimization problem is formulated
in gPROMS in which keeping the experimental conditions used by
Jin (2002), the initial particle size is set to the literature value and
then bounded by 710 times the initial value. The twodimensional form of the process model outlined in Section 2 is
used. The objective function to be maximized is dened for
methane conversion, Eq. (61), and overall H2 selectivity, Eq. (62),
which are investigated separately. The diameter and length of the
cylindrical pellet are treated as time-invariant variables to be
optimized.
F CH4 ;in  F CH4 ;out
F CH4 ;in

Table 13
Effect of particle size on X CH4 and SH2 for CLR xed-bed setup reported by Jin (2002)
at 600 1C.

performance. A smaller particle diameter should be selected to limit


the effect of internal transport resistance on reaction rates and
increase CH4 conversion and H2 selectivity. A decrease from
4.0 mm to 0.8 mm in particle diameter brought signicant changes,
shown in Fig. 13. The optimal particle size produces a 12% increase in
CH4 conversion and 7.7% increase in H2 selectivity (Table 13). Further
decrease of Dp did not produce any measureable concentration
differences, indicating a homogeneous reaction regime. Furthermore,
overall improvements to the catalytic reaction rates are a result of
reducing the diffusion barrier. The rate of methane conversion is
enhanced and the overall reactions are more selective toward
hydrogen production. Intraparticle concentration and temperature
gradients are also minimized. The proposed particle size is realistic,
does not create issues in experimental operation, and results in
valuable benets in the reactor performance. It is worth mentioning
that after reducing the particle diameter, the pressure drop marginally increased from 0.0011 to 0.0086 bar. Close-to-isobaric conditions
are maintained for this particular experimental apparatus; however,
this should not be expected for a larger-scale commercial process.

3.5. Optimization of particle size

X CH4

Fig. 13. Heterogeneous model predictions for the effect of particle size on the CLR
xed-bed experiment conducted by Jin (2002). Solid lines represent the reference
particle and dashed lines represent the optimal particle.

Model results reveal that the reaction rates are highly sensitive to
particle diameter, and insensitive to length. Transport resistance is
mainly associated with radial diffusion within the particle due to
good mixing in the axial direction. Therefore, a range of particle
lengths can be utilized without decreasing the oxygen carrier

The dynamic behavior of a xed-bed reactor for chemicallooping reduction is examined by a heterogeneous model, which is
shown to agree very well with literature data, for a variety of Nibased oxygen carriers, operating conditions, and reaction temperatures. As large particles are found to be affected by intra- and

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

inter- diffusion limitations, more accurate predictions are obtained


using the heterogeneous model. Differences between kinetic
parameters estimated from the homogeneous model and the
heterogeneous model are attributed to the extent of diffusion
limitations in the experiments. For large oxygen carriers that are
millimeters in size, diffusion limitations on some of the catalytic
reactions are found to be severe. Furthermore, temperature
uctuations in the xed-bed reactor are minimized by high rates
of dispersion and small reactor scales. While intraparticle temperature gradients are found to be negligible for most cases,
interphase heat exchange can be a limiting factor and should be
accounted for in modeling. Application of the heterogeneous
model is extended to optimize the particle size in order to improve
chemical-looping performance. Through reduction of the particle
diameter, diffusion-limited barriers are relaxed and as a result, CH4
conversion and H2 selectivity are enhanced. This reactor model,
which combines particle and uid effects, should enhance the
design of chemical-looping in xed-bed reactors and pave the way
for feasibility analysis of the chemical-looping technology. In Part
II of this work, a particle model will be used to analyze experiments performed in thermogravimetric analyzers. Physical and
chemical changes incurring to the Ni-based oxygen carrier due to
interaction with Al2O3 and from prolonged exposure to CLC
conditions will be discussed.
Nomenclature
Ac
Aij
al
a0
av
Ci
Cc,i
CNi
CNiO
CNiO
Cp,c
Cp,f
Cp,i
Cp,s
CT
CT,c
Dax,i
De,i
Dij
DK,i
Drad,i
Dm,i
Dp
Dt
Fi
Fin
FT
Fv,i
G
hf
hbed
hw
i

cross-sectional area of the reactor tube [m2]


binary interaction parameter
characteristic length of particle (Vp/Sp)
initial specic surface area of the oxygen carrier [m2/
kg OC]
external particle surface area per unit volume reactor
volume [1/m]
concentration of gas species i in uid phase [mol/m3]
concentration of gas species i in solid phase [mol/m3]
NiO concentration [kg Ni/kg OC]
Ni concentration [kg Ni/kg OC]
initial NiO concentration [kg Ni/kg OC]
heat capacity of the gas mixture of the solid phase
[J/mol K]
heat capacity of the gas mixture of the uid phase
[J/mol K]
heat capacity of the gas species [J/mol K]
heat capacity of the oxygen carrier [J/mol K]
total gas concentration in uid phase [mol/m3]
total gas concentration in solid phase [mol/m3]
axial dispersion coefcient of species i [m2/s]
effective diffusion coefcient of species i [m2/s]
binary gas phase diffusivities [cm2/s]
Knudsen diffusion coefcient of species i [m2/s]
radial dispersion coefcient of species i [m2/s]
diffusion coefcient of component i in the mixture [m2/s]
diameter of oxygen carrier [m]
diameter of the reactor tube [m]
molar ow rate of gas species i [mol/s]
molar ow rate of gas species i at the inlet [mol/s]
total molar ow rate of gas species [mol/s]
volume fraction of solid phase i
mass ux of the gas phase [kg/m2/s]
heat transfer coefcient between bulk uid and oxygen
carrier particle [W/m2 K]
packed-bed heat transfer coefcient [W/m2 K]
wall heat transfer coefcient [W/m2 K]
gas phase species (CH4, H2, H2O, CO, CO2, Ar, N2)

j
kc,i
KCO,j
K CO2;j
K CH4;j
K H2;j
K5
K6
K7
K8
K9
K10
k1k4
k11k41
k5
k6
k7
k8
k9
k10
k51k101
L
Lp
Mi
Nu
OC
P
Pi
Pc
Pea
Per
r
rc
rj
R
Rp
Re
Rei
Rep
Sc
Sg
Sh
Sp
t
T
Tc
Tin
Tr
Ts
Tw
ui
u
u0
U
V
Vp
X

247

chemical reaction
mass transfer coefcient between bulk uid and oxygen
carrier particle [m/s]
adsorption coefcient of CO [bar  1]
adsorption coefcient of CO2 [bar]
adsorption coefcient of CH4 [bar1]
adsorption coefcient of H2 [bar1]
equilibrium constant for steam methane reforming [bar2]
equilibrium constant for water gas shift
equilibrium constant for dry reforming [bar2]
equilibrium constant for methane decomposition [bar]
equilibrium constant for carbon gasication by
steam [bar]
equilibrium constant for carbon gasication by CO2 [bar]
reaction rate constant for oxygen carrier reduction
reactions [m/s]
frequency factor for oxygen carrier reduction rate
constants
rate constant for steam methane reforming [bar0.5mol/
kgNi s]
rate constant for water gas shift [mol/bar kgNi s]
rate constant for dry reforming [mol/bar kgNi s]
rate constant for methane decomposition [bar-1]
rate constant for carbon gasication by steam [mol/
kgNi s]
rate constant for carbon gasication by CO2 [mol/kgNi-s]
frequency factor for catalytic reduction rate constants
bed length [m]
oxygen carrier particle length [m]
molar weight of species [g/mol]
uid to wall Nusselt number (hwDp/m)
oxygen carrier
reactor pressure [bar]
partial pressure of species i [bar]
critical pressure [bar]
axial Peclet number (u0Dp/Dax)
radial Peclet number (u0Dp/Drad)
reactor radial element
oxygen carrier radial element
rate of j reaction [mol/m3 s]
reactor radius [m]
oxygen carrier particle radius [m]
Reynolds number based on supercial velocity (uDp/)
Reynolds number based on interstitial velocity (uiDp/)
Reynolds number based on initial supercial velocity
(uoDp/(1 b)/)
Schmidt number (//Dm,i)
initial specic surface area of the oxygen carrier
[cm2/g OC]
Sherwood number (kc,iDp/Dm,i)
surface area of particle
time element [s]
temperature of the uid phase [1C]
critical temperature [1C]
temperature in the inlet [1C]
reduced temperature
temperature of the solid phase [1C]
temperature at the wall [1C]
interstitial velocity [m/s]
supercial velocity [m/s]
supercial velocity at the inlet [m/s]
overall heat transfer coefcient [J/m2 K s]
volume element
volume of particle
conversion of oxygen carrier

248

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

yi
yi,feed
z
zc

mole fraction of species i


mole fraction of species i in the feed
length element [m]
oxygen carrier axial element (cylindrical particle)

reduced inverse thermal conductivity


heat of reaction j [J/mol]
numerical constant, near unity
porosity of the oxygen carrier
porosity of the bed
RR
effectiveness factor for reaction Rj 3 0 p r 2c Rj dr c =R3p Rj 
axial heat dispersion coefcient [W/m K]
static contribution to effective thermal conductivity
[W/m K]
thermal conductivity of species i [W/m K]
radial heat dispersion coefcient [W/m K]
thermal conductivity of the gas mixture [W/m K]
thermal conductivity of the oxygen carrier [W/m K]
translational thermal conductivity for species i
viscosity of the gas mixture [N s/m2]
mass density of the gas [kg/m3]
bulk density of the xed-bed [kg/m3]
density of the oxygen carrier [kg/m3]
pore tortuosity

i
rad
m
s
tr,i

b
s

Acknowledgments
This material is based upon work supported by the National
Science Foundation under Grant No. 1054718.

Appendix
Mass and energy balances for transport through a cylindrical
particle are presented (Eqs. (63) and (64)).
 2

C
C c;i 1 C c;i 2 C c;i
c c;i De;i

63
p Ri
2
r c r c
t
zc
r c
c C T;c C p;c 1  c p C p;s

 2

T c
T c 1 T c 2 T c
s

2
2
t
r c r c r c zc

64

The boundary conditions using the cylindrical particles are


De;i

C c;i
j
kf ;i C c;i jrc Rc  C i
r c rc Rp

C c;i
j
0
r c rc 0

65
66

De;i

C p;i
j
kf ;i C p;i jzp Lp  C i
zp zp Lp

67

De;i

C c;i
j
kf ;i C c;i jzc Lp  C i
zc zc Lp

68

 s

T c
j
hf T p jrc Rp  T
r c rc Rp

T c
j
0
r c rc 0
 s

T c
j
hf T p jzc Lp  T
zc zc Lp

72

These equations can be combined with the uid phase heterogeneous model equations in either the one-dimensional or twodimensional form.

Greek letters

Hj

c
b
j
ax
e0

T c
j
hf T c jzc 0 T
zc zc 0

69
70
71

References
Aaron, D., Tsouris, C., 2005. Separation of CO2 from ue gas: a review. Separation
Science and Technology 40, 321348.
Abad, A., Mattisson, T., Lyngfelt, A., Rydn, M., 2006. Chemical-looping combustion
in a 300 W continuously operating reactor system using a manganese-based
oxygen carrier. Fuel 85, 11741185.
Adnez, J., Celaya, J., Diego, L.F., de, Garcia-Labiano, F., Abad, A., 2006. Chemical
looping combustion in a 10 kW th prototype using a CuO/Al2O3 oxygen carrier:
effect of operating conditions on methane combustion. Industrial & Engineering Chemistry Research 45, 60756080.
Adnez, J., Diego, L.F., de, Garcia-Labiano, F., Gayn, P., Abad, A., 2004. Selection of
oxygen carriers for chemical-looping combustion. Energy & Fuels 18, 371377.
Adnez, J., Dueso, C., Diego, L.F., de, Garcia-Labiano, F., Gayn, P., Abad, A., 2009.
Methane combustion in a 500 W th chemical-looping combustion system using
an impregnated Ni-based oxygen carrier. Energy & Fuels 23, 130142.
Aspen, 2007. Aspen Technology: Physcial Properties System.
Ahmet, K., Avci, Z., Ilsen, nsan., 2011. Reactor design for fuel processing, in:
Dushyant Shekhawat, James J. Spivey, David A. Berry, (Eds.) Fuel Cells:
Technologies for Fuel Processing, pp. 451516.
Bischoff, K.B., Levenspiel, O., 1962. Fluid dispersiongeneralization and comparison
of mathematical modelsII Comparison of models. Chemical Engineering
Science 17, 257264.
Davis, Mark E., Davis, Robert J., 2003. Fundamentals of Chemical Reaction
Engineering. 1st edition, McGraw-Hill, NY, p. 384.
De Diego, L.F., Ortiz, M., Garca-Labiano, F., Adnez, J., Abad, A., Gayn, P., 2009.
Hydrogen production by chemical-looping reforming in a circulating uidized
bed reactor using Ni-based oxygen carriers. Journal of Power Sources 192,
2734.
Dueso, C., Ortiz, M., Abad, A., Garca-Labiano, F., De Diego, L.F., Gayn, P., Adnez, J.,
2012. Reduction and oxidation kinetics of nickel-based oxygen-carriers for
chemical-looping combustion and chemical-looping reforming. Chemical Engineering Journal 188, 142154.
Fan, L.-S., 2011. Chemical Looping Systems for Fossil Energy Conversions. John
Wiley & Sons.
Finkenrath, M., 2011. International Energy Agency: Cost and Performance of Carbon
Dioxide Capture from Power Generation.
Fogler, H.S., 2005. Elements of Chemical Reaction Engineering, 4th ed.
Foumeny, E.A., Chowdhury, M.A., McGreavy, C., Castro, J.A.A., 1992. Estimation of
Dispersion Coefcients in Packed Beds. Chemical Engineering & Techology 15,
161181.
Garcia-Labiano, F., Adnez, J., Diego, L.F., de, Gayan, P., Abad, A., 2006. Effect of
pressure on the behavior of copper-, iron-, and nickel-based oxygen carriers for
chemical-looping combustion. Energy & Fuels 20, 2633.
Garca-Labiano, F., De Diego, L.F., Adnez, J., Abad, A., Gayn, P., 2005. Temperature
variations in the oxygen carrier particles during their reduction and oxidation
in a chemical-looping combustion system. Chemical Engineering Science 60,
851862.
Halmann, M., Steinberg, M., 2000. Greenhouse Gas Carbon Dioxide Mitigation:
Science and Technology. Lewis Publishers, Boca Raton, Fl.
Hossain, M.M., De Lasa, H.I., 2008. Chemical-looping combustion (CLC) for inherent
separationsa review. Chemical Engineering Science 63, 44334451.
Iliuta, I., Tahoces, R., Patience, G.S., Rifart, S., Luck, F., 2010. Chemical-looping
combustion process: kinetics and mathematical modeling. AIChE Journal 56,
10631079.
IPCC, 2005. IPCC Special Report on Carbon Dioxide Capture and Storage.
IPCC, 2007a. IPCCIntergovernmental Panel on Climate Change.
IPCC, 2007b. IPCC Fourth Assessment Report.
IPCC, 2008. IPCCIntergovernmental Panel on Climate Change. Cambridge, U.K.
Ishida, M., Jin, H., 1996. A novel chemical-looping combustor without NOx
formation. Industrial & Engineering Chemistry Research 35, 24692472.
Ishida, M., Jin, H., Okamoto, T., 1996. A fundamental study of a new kind of medium
material for chemical-looping combustion. Energy & Fuels 10, 958963.
Ishida, M., Zheng, D., Akehata, T., 1987. Evaluation of a chemical-looping combustion power-generation system by graphic energy analysis. Energy 12, 147154.
Jakobsen, H.A., 2011. Fixed bed reactors, University of Science and Technology, in
TKP4115 Reactor Technology, Spring. http://www.nt.ntnu.no/users/jakobsen/
TKP4145/xedbed_2011.pdf.
Jerndal, E., Mattisson, T., Thijs, I., Snijkers, F., Lyngfelt, A., 2009. NiO particles with
Ca and Mg based additives produced by spray- drying as oxygen carriers for
chemical-looping combustion. Energy Procedia 1, 479486.
Jin, H., 2002. Reactivity study on natural-gas-fueled chemical-looping combustion
by a xed-bed reactor. Industrial & Engineering Chemistry Research 41,
40044007.

L. Han et al. / Chemical Engineering Science 104 (2013) 233249

Jin, H., Okamoto, T., Ishida, M., 1999. Development of a novel chemical-looping
combustion: synthesis of a solid looping material of NiO/NiAl2O4. Industrial &
Engineering Chemistry Research 38, 126132.
Johansson, E., Mattisson, T., Lyngfelt, A., Thunman, H., 2006a. A 300 W laboratory
reactor system for chemical-looping combustion with particle circulation. Fuel
85, 14281438.
Johansson, E., Mattisson, T., Lyngfelt, A., Thunman, H., 2006b. Combustion of syngas
and natural gas in a 300 W chemical-looping combustor. Chemical Engineering
Research and Design 84, 819827.
Johansson, M., Mattisson, T., Rydn, M., Lyngfelt, A., 2006c. Carbon capture via
chemical-looping combustion and reforming chemical-looping combustion.
In: International Seminar on Carbon Sequestration and Climate Change. Rio
de Janeiro, Brazil, pp. 2427.
Kerr, H.R., 2005. Capture and separation technologies gaps and priority research
needs. In: Carbon Dioxide Capture for Storage in Deep Geologic Formations
Result from the CO2 Capture Project.
Kronberger, B., Johansson, E., Lofer, G., Mattisson, T., Lyngfelt, A., Hofbauer, H.,
2004. A two-compartment uidized bed reactor for CO2 capture by chemicallooping combustion. Chemical Engineering & Technology 27, 13181326.
Kronberger, B., Lyngfelt, A., Lofer, G., Hofbauer, H., 2005. Design and uid dynamic
analysis of a bench-scale combustion system with CO2 separation-chemicallooping combustion. Industrial & Engineering Chemistry Research 44, 546556.
Linderholm, C., Abad, A., Mattisson, T., Lyngfelt, A., 2008. 160 h of chemical-looping
combustion in a 10 kW reactor system with a NiO-based oxygen carrier.
International Journal of Greenhouse Gas Control 2, 520530.
Linderholm, C., Jerndal, E., Mattisson, T., Lyngfelt, A., 2010. Investigation of NiObased mixed oxides in a 300-W chemical-looping combustor. Chemical
Engineering Research and Design 88, 661672.
Linderholm, C., Mattisson, T., Lyngfelt, A., 2009. Long-term integrity testing of
spray-dried particles in a 10-kW chemical-looping combustor using natural gas
as fuel. Fuel 88, 20832096.
Lyngfelt, A., Leckner, B., Mattisson, T., 2001. A uidized-bed combustion process
with inherent CO2 separation; application of chemical-looping combustion.
Chemical Engineering Science 56, 31013113.
Marrero, T.R., Mason, E.A., 1972. Gaseous diffusion constants. Journal of Physical
and Chemical Reference Data 1, 1118.
Mattisson, T., Johansson, M., Jerndal, E., Lyngfelt, A., 2008. The reaction of NiO/
NiAl2O4 particles with alternating methane and oxygen. The Canadian Journal
of Chemical Engineering 86, 756767.
Mattisson, T., Johansson, M., Lyngfelt, A., 2006. The use of NiO as an oxygen carrier
in chemical-looping combustion. Fuel 85, 736747.
Mendiara, T., Johansen, J.M., Jensen, A.D., Glarborg, P., 2011. Evaluation of different
oxygen carriers for biomass tar reforming (II): carbon deposition in experiments with methane and other gases. Fuel 90, 13701382.
Noorman, S., Gallucci, F., Van Sint Annaland, M., Kuipers, J. a.M., 2011a. A theoretical
investigation of CLC in packed beds. Part 1: Particle model. Chemical Engineering Journal 167, 297307.
Noorman, S., Gallucci, F., Van Sint Annaland, M., Kuipers, J. a.M., 2011b. A theoretical
investigation of CLC in packed beds. Part 2: Reactor model. Chemical Engineering Journal 167, 369376.
Noorman, S., Van Sint Annaland, Martin, Kuipers, H., 2007. Packed bed reactor
technology for chemical-looping combustion. Industrial & Engineering Chemistry Research 46, 42124220.
Ortiz, M., Diego, L.F., De, Abad, A., Garca-labiano, F., Gaya, P., Ada, J., 2012. Catalytic
activity of ni-based oxygen-carriers for steam methane reforming in chemicallooping processes. Energy & Fuels 26, 791800.
Perry, R.H., Green, D.W., 1997. Perry's Chemical Engineer's Handbook, 7th ed.

249

Peters, P.E., Schifno, R.S., Harriott, P., 1988. Heat transfer in packed-tube reactors.
Industrial & Engineering Chemistry Research 27, 226233.
Poling, B.E., Prausnitz, J.M., O'Connel, J.P., 2000. The Properties of Liquids and Gases,
5 edition. McGraw-Hill Professional.
Process Systems Enterprise Limited, 2012.
Pushnov, a.S., 2006. Calculation of average bed porosity. Chemical and Petroleum
Engineering 42, 1417.
Richter, H.J., Knoche, K.F., 1983. Reversibility of combustion processes. In: ACS
Symposium Series, pp. 7185.
Rydn, M., Lyngfelt, A., Mattisson, T., 2006. Synthesis gas generation by chemicallooping reforming in a continuously operating laboratory reactor. Fuel 85,
16311641.
Rydn, M., Lyngfelt, A., Mattisson, T., 2008a. Chemical-looping combustion and
chemical-looping reforming in a circulating uidized-bed reactor using Nibased oxygen carriers. Energy & Fuels 22, 25852597.
Rydn, M., Lyngfelt, A., Mattisson, T., Chen, D., Holmen, A., Bjrgum, E., 2008b.
Novel oxygen-carrier materials for chemical-looping combustion and chemicallooping reforming; LaxSr1  xFeyCo1  yO3  perovskites and mixed-metal oxides
of NiO, Fe2O3 and Mn3O4. International Journal of Greenhouse Gas Control 2,
2136.
Ryu, H.-J., Bae, D.-H., Jin, G.-T., 2003. Effect of temperature on reduction reactivity of
oxygen carrier particles in a xed bed chemical-looping combustor. Korean
Journal of Chemical Engineering 20, 960966.
Sahir, A.H., Lighty, J.S., Sohn, H.Y., 2011. Kinetics of copper oxidation in the air
reactor of a chemical looping combustion system using the law of additive
reaction times. Industrial & Engineering Chemistry Research 50, 1333013339.
Sattereld, C.N., 1981. Mass Transfer in Heterogeneous Catalysis, 2nd ed..
Sedor, K.E., Hossain, M.M., De Lasa, H.I., 2008. Reactivity and stability of Ni/Al2O3
oxygen carrier for chemical-looping combustion (CLC). Chemical Engineering
Science 63, 29943007.
Shulman, A., Linderholm, C., Mattisson, T., Lyngfelt, A., 2009. High reactivity and
mechanical durability of NiO/NiAl2O4 and NiO/NiAl2O4/MgAl2O4 oxygen carrier
particles used for more than 1000 h in a 10 kW CLC reactor. Industrial &
Engineering Chemistry Research 48, 74007405.
Son, S.R., Kim, S.D., 2006. Chemical-looping combustion with NiO and Fe2O3 in a
thermobalance and circulating uidized bed reactor with double loops.
Industrial & Engineering Chemistry Research 45, 26892696.
Wako, N., Kaguei, S., Funazkri, T., 1979. Effect of uid dispersion coefcients on
particle-to-uid heat transfer coefcients in packed beds: correlation of nusselt
numbers. Chemical Engineering Science 34, 325336.
Wilkes, J.O., 1999. Fluid Mechanics for Chemical Engineers.
Yagi, S., Kunii, D., Wakao, N., 1960. Studies on axial effective thermal conductivities
in packed beds. AIChE Journal 6, 543546.
Yagi, S., Wakao, N., 1959. Heat and mass transfer from wall to uid in packed beds.
AIChE Journal 5, 7985.
Zafar, Q., Mattisson, T., 2005. Integrated hydrogen and power production with CO2
capture using chemical-looping reformings redox reactivity of particles of CuO,
Mn2O3, NiO, and Fe2O3 using SiO2 as a support. Industrial & Engineering
Chemistry Research 44, 34853496.
Zafar, Q., Mattisson, T., Gevert, B., 2006. Redox investigation of some oxides of
transition-state metals Ni, Cu, Fe, and Mn supported on SiO2 and MgAl2O4.
Energy & Fuels 20, 3444.
Zhou, Z., Han, L., Bollas, G.M. Overview of chemical-looping reduction in xed-bed
and uidized-bed reactors focused on oxygen carrier utilization and reactor
efciency. Aerosol and Air Quality Research, submitted for publication.
Zhou, Z., Han, L., Bollas, G.M., 2013. Model-based analysis of bench-scale xed-bed
units for chemical-looping combustion. Chemical Engineering Journal 233,
331348.

Das könnte Ihnen auch gefallen