You are on page 1of 17

Fuel 121 (2014) 327–343

Contents lists available at ScienceDirect

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

Design of a rotary reactor for chemical-looping combustion.


Part 1: Fundamentals and design methodology
Zhenlong Zhao, Chukwunwike O. Iloeje, Tianjiao Chen, Ahmed F. Ghoniem ⇑
Department of Mechanical Engineering, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139-4307, 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 novel rotary reactor has been


proposed for CLC with carbon
separation.
 The design is compact, stable, and it
captures CO2 with minimum energy
penalty.
 The fundamentals of OC, reactor
design, and operating conditions have
been studied.
 The key parameters for the reaction
kinetics have been evaluated by
simple models.
 A systematic procedure has been
proposed for reactor design and
optimization.

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 novel and promising option for several applications including
Received 18 August 2013 carbon capture (CC), fuel reforming, H2 generation, etc. Previous studies demonstrated the feasibility
Received in revised form 22 November 2013 of performing CLC in a novel rotary design with micro-channel structures. In the reactor, a solid wheel
Accepted 25 November 2013
rotates between the fuel and air streams at the reactor inlet, and depleted air and product streams at exit.
Available online 11 December 2013
The rotary wheel consists of a large number of micro-channels with oxygen carriers (OC) coated on the
inner surface of the channel walls. In the CC application, the OC oxidizes the fuel while the channel is in
Keywords:
the fuel zone to generate undiluted CO2, and is regenerated while the channel is in the air zone. In this
Chemical-looping combustion
Rotary reactor
two-part series, the effect of the reactor design parameters is evaluated and its performance with differ-
CO2 capture ent OCs is compared. In Part 1, the design objectives and criteria are specified and the key parameters
Oxygen carriers controlling the reactor performance are identified. The fundamental effects of the OC characteristics,
Reaction kinetics the design parameters, and the operating conditions are studied. The design procedures are presented
on the basis of the relative importance of each parameter, enabling a systematic methodology of selecting
the design parameters and the operating conditions with different OCs. Part 2 presents the application of
the methodology to the designs with the three commonly used OCs, i.e., nickel, copper, and iron, and
compares the simulated performances of the designs.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction atmosphere and global warming. One approach to reduce anthro-


pogenic CO2 emissions, apart from improving energy efficiency
It has been widely accepted that fossil fuel combustion is a and using alternative sustainable energy sources, is carbon capture
major contributor to the rise of CO2 concentration in the and sequestration (CCS), in which CO2 is separated from flue gases,
liquefied, and injected in geological formations, such as depleted
oil or gas fields. Of these steps, CO2 capture remains the most
⇑ Corresponding author. Tel.: +1 617 253 2295.
challenging part from an economic point of view since obtaining
E-mail address: ghoniem@mit.edu (A.F. Ghoniem).

0016-2361/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.fuel.2013.11.056
328 Z. Zhao et al. / Fuel 121 (2014) 327–343

Nomenclature
X conversion of oxygen carrier
Symbols xi molar fraction of species i
A cross-sectional area, m2 z99% height to reach 99% fuel conversion, m
Ci concentration of species i, mol m3
Cf friction coefficient Greek letters
c specific heat capacity, J kg1 K1 dbulk thickness of the bulk support layer, m
D reactor diameter, m doc thickness of the porous oxygen carrier layer, m
dh channel width, m e porosity of the OC
H channel height, m gI fuel conversion efficiency, percentage of fuel converted
hgs heat transfer coefficient between phases, W m2 K1 in the reactor
hm external mass transfer coefficient, m s1 cCH4 conversion yield of CH4
Dhr enthalpy of reaction, J mol1 j redundancy factor
k reaction rate constant lg viscosity of gas, kg m1 s1
ks,kg thermal conductivity of solid or gas phase, W m1 K1 m stoichiometric coefficient
k0 pre-exponential factor hi size of sector i, rad
mox mass of fully oxidized oxygen carrier, kg q density, kg m3
mred mass of fully reduced oxygen carrier, kg s cycle period, characteristic time of reaction, channel
N_ f molar flow rate of fuel residence time, s
P operating pressure, Pa x rotational velocity, s1
Pc inner perimeter of the channel, m
pi partial pressure of species i, Pa Acronyms
Roc oxygen transport capacity CC carbon capture
Ru gas universal constant, 8.314 J mol1 K1 CCS carbon capture and sequestration
T temperature, K CLC chemical-looping combustion
t gas residence time in the channel, s1 OC oxygen carrier
u velocity, m s1 Redox reduction and oxidation
W _ th target thermal capacity, W TIT turbine inlet temperature

carbon dioxide in high purity still accounts for the major share of consists of a rotary wheel and two stationary chambers at the top
the cost of state-of-the-art CCS technologies. In the past few years, and bottom of the wheel. The wheel rotates continuously through
extensive research focus has been placed on three general pro- four sectors (Fig. 1b): fuel, air, and two purging sectors. The rotary
cesses for capturing CO2 from combustion in power plants: post- wheel consists of a large number of micro-channels (Fig. 1c) with
combustion capture, pre-combustion decarbonization and oxy- the OC coated onto their inner wall. As shown in Fig. 1d, the chan-
combustion. One of the key issues that limits the applications of nel wall has two solid layers with one being a highly porous OC
CCS approaches is the large energy penalty during the separation layer and the other being a bulk dense ceramic layer with high
process. thermal inertia and thermal conductivity. The OC layer usually
Recently, a new approach for CO2 capture has been widely consists of an active metal oxide which reacts with the fuel and
investigated. This approach was named ‘‘chemical-looping com- air alternatively, as well as an inert ceramic substrate which helps
bustion (CLC)’’ [1] and belongs to oxy-fuel combustion. In CLC, maintain the pore structures and chemical reactivities of the OC.
combustion is decomposed into two steps: fuel is oxidized by a As seen in Fig. 2, pressurized feed gas (fuel, air or steam) flows
metal oxide in a fuel reactor to generate CO2 and steam; the re- through the reactor, reacts with the OC while it is heated to high
duced metal oxide is then regenerated by air in an air reactor. Dur- temperature. In the fuel sector, the fuel is diluted with CO2 to
ing this two-step process, the looping medium acts as an ‘‘oxygen effectively lower the operating temperature while in the air sec-
carrier’’ (OC), which adsorbs oxygen in the air reactor and releases tor air is used to re-oxidize the OC. Two purging sectors are
it to oxidize fuel in the fuel reactor. The main advantage of CLC is implemented between the fuel and the air sectors to sweep the
that by using OCs as the medium to transport the pure oxygen, the residual gases out of the reactor. Flue streams from a large num-
direct contact between air and fuel is circumvented and hence en- ber of channels merge into two separate streams from the fuel
ergy-intensive gas separation processes are avoided. zone and the air zone (see Fig. 2b), respectively, and then drive
CLC is commonly carried out by physically transporting the OC turbines in the downstream, as described in Ref. [22]. The main
particles between two interconnected fluidized bed reactors [2–4]. incentives for performing CLC in this way are (1) the separation
In this reactor concept, the OCs in the form of particles are fluidized of the gas and particles is intrinsically avoided, (2) the operation
and pneumatically transported continuously between the fuel and is continuous and stationary, and (3) the design is compact and
air reactors. Several CLC units using the fluidized-bed design have easy to scale.
been built and operated by Chalmers University of Technology of Pavone and co-workers [24,25] simulated the reduction and
Sweden [5–8], ICB-CSIC of Spain [9–11], Vienna University of Tech- oxidation (redox) performances of the rotary reactor with nickel
nology of Austria [12,13], and Southeast University of China oxide wash-coated on alumina substrate using COMSOL for the
[14,15]. Meanwhile, some novel reactor designs have emerged initial cycles. Because of the limited amount of the solid phase in
such as the moving-bed reactor in Ohio State University [16,17], the base-case design, large temperature fluctuations (500 °C) were
the fixed packed-bed reactor in Eindhoven University of Technol- observed in the solid phase, which renders the design unstable
ogy of Netherlands [18,19], and the rotating packed-bed reactor over repeated cycles. Zhao et al. [26] investigated the periodic
in SINTEF Materials & Chemistry of Norway [20,21]. stationary-state performances of a rotary design with copper oxide
Recently, a rotary reactor design with micro-channel supported on boron nitride substrate using a one-dimensional
structures was proposed [22,23]. As shown in Fig. 1a, the reactor plug-flow model. The base-case simulation validated the
Z. Zhao et al. / Fuel 121 (2014) 327–343 329

Fig. 1. Schematic diagram of the rotary CLC design [22]: (a) front view, (b) bottom view, (c) individual channel structure, and (d) the OC coated on the surface.

Fig. 2. Schematic illustration of the reactor with inlet and exit flue streams. Panel (a) is the isometric projection of the reactor, and panels (b) and (c) are the gas composition
at the exit and the inlet of the reactor for two cycles.

applicability of the rotary design to the CLC process with high The aim of this two-part series is to investigate the fundamental
separation efficiency and operational stability. However, the sensi- effects of design on the rotary reactor performance, and to propose
tivity study showed that the performances are highly dependent a comprehensive methodology of selecting the suitable materials
on the selection of the OCs, the design of the reactor, and the and the operating conditions. In Part 1, the design objectives and
operating conditions. These parameters are highly coupled, and criteria are discussed. The fundamental effects of the OC character-
they are all closely related to the performance of the reactor. istics, the design parameters, and the operating conditions are
Therefore, the fundamentals of such a system and effects of design studied. The design procedures are presented on the basis of the
parameters must be studied thoroughly and systematic relative importance of each parameter, enabling a systematic
approaches and methodologies of selecting the design and operat- methodology of selecting the design parameters and the operating
ing conditions should be investigated in depth. conditions with different OCs. Part 2 presents the application of the
methodology to the designs with three commonly used OCs, i.e.,
330 Z. Zhao et al. / Fuel 121 (2014) 327–343

nickel, copper, and iron, and compares the simulated performances larger number of active sites and enhancing the contact between
of the designs. the gas and solid phases; thin grain layers (dgrain) with well-dis-
persed active metal oxides on the support effectively reduce the
2. Design objectives and criteria product layer diffusion resistance through the grain and thus en-
hance the exposure of the embedded ions to the gaseous reactants.
The design of the rotary CLC reactor should satisfy a number of For an optimized design, the rule of thumb is to minimize the resis-
requirements, including (1) complete fuel conversion, (2) CO2 sep- tance from each step and maintain Dai below 0.1. As a result, the
aration, and (3) operational stabilities during repeated cycles. In reaction should be mainly controlled by the intrinsic chemical
addition, the reactor design should also be (4) compatible and easy kinetics of the selected metal oxide.
to integrate with the power cycle, and (5) economically feasible To further raise the fuel conversion rate, favorable operating
with low fabrication cost. This section addresses the objectives conditions can be selected on the basis of the OCs. Higher fuel con-
and the criteria of design and also identifies the key parameters centration or temperature always accelerates the reaction rate,
that control the reactor performance. although the dependency varies significantly among different
OCs. For example, as noted in Part 2, the reduction rate of the
2.1. Fuel conversion iron-based OC is mostly sensitive to the concentration of the fuel
while the high endothermicity of nickel oxide reduction leads to
High fuel conversion is required for the reactor design. Fuel con- a high reaction barrier and therefore the rate correlates with the
version is controlled by the reduction reaction between the gas- temperature. Operating pressures (P) may also affect the reaction
eous fuel and the active metal oxide on the surface. For this rate because the gas composition, the species diffusion, the fuel
purpose, one needs to (1) use an OC layer with highly reactive me- adsorption, the morphology and topology of the grain, and the
tal oxides, (2) operate at favorable conditions (Tin, P, etc.), or (3) amount of the open sites on the grain may change with pressure.
provide sufficient residence time (tfuel) of gas. The degree of OC conversion also has influences on the reduction
As seen in Fig. 3, the reduction reaction within the reactor gen- rate, and generally the reaction rate is highest initially, and slows
erally involves several key steps: external and internal diffusions, down as the conversion proceeds. Thus, an intermediate OC con-
adsorption and dissociation, diffusion through the grain, and reac- version range is usually preferred, where the OC is neither fully
tions. Table 1 summarizes the characteristic times for the conver- oxidized nor fully reduced [17].
sion of the OC under each resistance (see Section 3.1.1 for Adequate residence time of the gas in the channel (tfuel) should
derivation). A highly reactive metal oxide, e.g., nickel-, copper- be maintained to ensure complete fuel consumption:
oxide, with a suitable oxidation state, e.g., Fe2O3/Fe3O4, favors the Z H
dz
conversion efficiency. In addition, properly selected OC preparation tfuel ¼ P t 99%;fuel ð1Þ
0 ufuel
methods and layer designs minimize the resistance from each
intermediate step. For instance, as seen in Table 1, smaller Here, t99%,fuel is the characteristic time of 99% fuel conversion, ufuel is
channels with a higher convective transport rate (hm) reduces the the bulk velocity in the fuel sector, and H is the height of the reactor
external resistance; a higher porosity increases the effective channels. To increase the residence time (tfuel), one can use longer
diffusivity (De) through the pores and thus enhances internal diffu- channels (H) or a lower velocity (ufuel), both of which, however, re-
sion; a highly porous OC layer with large specific areas (soc) quire a larger reactor to reach the target thermal output, and hence
decreases the resistances from steps 3–5 (Table 1) by providing a increase the fabrication cost. On the other hand, the characteristic
time (t99%,fuel) for conversion can be reduced by admitting the same
amount fuel at the inlet with a lower flow rate but a higher concen-
tration (xfuel). Thus, the high fuel concentration favors the reactivity
and hence enhances the conversion (see Part 2). However, this effect
may be weakened when most of the OC near the inlet is consumed by
fuel, leading to a decreased reaction rate near the inlet and thus an
gradually increased t99%,fuel. Besides, a higher air flow rate (uair) is
required to effectively transfer the reaction heat by convection, since
in this case less dilute gas (CO2) is fed in the fuel sector.

2.2. Carbon separation

To capture CO2 generated in the fuel sector, the residual gas in-
side the channel should be purged within the fuel purge sector.
Similarly, the residual air inside the channel should be removed
as the channel passes through the air purge sector. Thus, the design
requires that
Z H
dz hi
ti ¼ 6 si ¼ i ¼ fuel purge or air purge ð2Þ
0 ui x
ti being the residence time of gas, ui the purge velocity, si the resi-
dence time of the channel in the (fuel or air) purge sector, hi the
purge sector size and x the rotational velocity. To reduce ti, shorter
channels (H) and higher purge velocities (ui) are preferred in both
purge sectors. However, a sufficient tfuel purge should be adapted,
i.e., tfuel,purge P t99%,fuel purge, to convert the residual fuel within the
channel, similar as in the fuel sector (Eq. (1)). In contrast, the
Fig. 3. Schematic drawing of the reactive flow in the channel. steam velocity in the air purge sector can be much higher and hence
Z. Zhao et al. / Fuel 121 (2014) 327–343 331

Table 1
Algebra expressions for the characteristic time of conversion under the control of each resistance.

# Resistance Conversion equations Characteristic time of conversion Damköhler number


1 External diffusion to the OC surface XB = t/sex sex ¼ mq^r hBmdocC A Daex ¼ soc qhsmkr doc
~

2 Internal diffusion through pores X 2B ¼ t=sin


2
sin ¼ 2mq^rBCdAocDe s q ~ s kr d2oc
Dain ¼ oc 2D e

3 Adsorption and dissociation XB = t/sad sad ¼ v r soc qq~^sBkad CA kr


Daad ¼ k
ad
q~ s kr dgrain
4 Diffusion in the product layer of a grain X 2B ¼ t=sd;g sd;g ¼ 2mqr sBocdCgrain
^
A De;g
Dad;g ¼ 2De;g
5 Chemical reaction in the grain ln(1  XB) = t/sr sr ¼ v r socq^q~Bs kr CA –

tair,purge can be greatly decreased. Thus, the fuel purge sector gener- Extensive experiments on nickel-, copper-, and iron-based me-
ally takes a larger share of the reactor than the air purge sector. On tal oxides have illustrated their long-term repeatability as OCs for
the other hand, one can improve the separation efficiency by using a CLC, while manganese-, cobalt-, or calcium sulfate-based OCs may
slower rotational velocity (x). A slower rotational velocity (x) in- decompose under high temperature or with the presence of CO2
creases the residence time of each channel in the purging sector and steam in the reactor [4]. The inert substrate in the porous layer
(si). Therefore, more purging steam flushes through the reactor to should act as a structure support to improve the gas permeability,
clean up the residual gas (CO2, or O2/N2). The decrease of x, how- enhance the active surface areas, sustain the stresses during cyclic
ever, is constrained by the concerns of fuel conversion and the operations, and hence improve the chemical stability of the OCs.
thermal stability of the reactor [26,27]: (1) at a slower x, and con- Ceramics (e.g., Al2O3, YSZ) are commonly suggested as the proper
sequently a longer residence time of channel within the fuel sector binder materials because of their excellent corrosion resistance un-
(sfuel), more OCs are consumed by fuel, leading to a reduced reduc- der high-temperature oxidizing environment. The support material
tion rate, and hence potentially a lower fuel conversion efficiency; should also be compatible to the selected metal oxide.
(2) more chemical energy is released from the redox reactions dur- In addition, the OC should be sufficiently resistant to carbon
ing one cycle, leading to a larger temperature fluctuation. deposition, melting and agglomeration, and impurity species
In addition, the fuel leakage and the air dilution between the fuel (e.g., sulfur) in the feed gas. As discussed before, carbon formation
and air zones will decrease the separation efficiency. In the rotary risk can be mitigated by selecting less catalyzing materials, adding
reactor, gas leakage may originate from the pressure difference be- CO2 and/or steam, or maintaining sufficient oxygen within the OCs
tween adjacent sectors, as well as the spinning motion of the rotary during the fuel sector. The melting of the materials (e.g., metal
wheel [22]. To minimize the pressure-driven gas leakage, relatively oxide, support substrate, or impurities) within the porous layer
uniform operating pressure (P) is recommended in the system may lead to irreversible changes of the pore structure. The impuri-
among different sectors. In addition, the pressure drop along the ties, such as H2S, within the gaseous fuel may react with the active
channel caused by skin friction should be similar among different metal oxide to form compounds that are poisonous to the OCs, de-
channels in order to match the pressure distribution in both feed crease the OC capacity, block the pore structure, and significantly
and exit chambers. Therefore, the similar feed velocities are recom- impairs the stability, and thus it is generally preferred to operate
mended among different sectors, while steady conversion patterns under an impurity-free environment [4].
are preferred to minimize the pressure field variation within the The thermal stability requires the reactor be resistant to the
reactor. Effective sealing systems (brush seals, labyrinth seals [28]) thermal stresses during the repeated cycles. The thermal expan-
from gas turbines can be adapted to the rotary reactor to limit the sions induced instabilities in the reactor primarily originate from
spinning-driven dilution between the fuel and the air zones. the mismatch of the support materials in the two-layer structure,
Carbon deposition on the channel walls during the fuel sector the axial temperature variation, and the peripheral temperature
may be transported to the following air sector, leading to reduced fluctuation:
separation efficiency. The deposition of carbon mainly depends on
the type of the material and the availability of oxygen [4]. Metal  Similar support materials with high toughness but low coeffi-
catalysts may catalyze the fuel decomposition reaction, and lead cient of thermal expansion should be used to match the expan-
to the formation of carbon. This appears more prominent near sion of the porous and the dense layers, and hence reduce the
the end of the reduction period as the majority of the OC is con- thermal distortion and stresses between the two layers.
sumed [29]. On the other hand, deposited carbon can be eliminated  In the periodic operation, the solid temperature increases from
via gasification with H2O or CO2 [30–34] or the solid–solid reaction the inlet to the exit. The axial temperature variations are
between carbon and the lattice oxygen in the OC [35]. Therefore, accompanied by differential thermal expansion. Thus, sufficient
there is usually a balance between carbon formation and disap- clearances should be maintained between the rotary wheel and
pearance, and thermodynamic equilibrium usually determines the outer insulating walls to account for the expansion. To min-
the relative importance of them. Adding sufficient H2O (or CO2) imize the differential expansion along the channel, a thick and
[36] and maintaining sufficient oxygen within the OC [22] shift highly conductive dense support layer is preferred to effectively
the equilibrium and hence mitigate the carbon formation risk. transfer the heat from the exit to the inlet.
 The local solid temperature may vary with time as the channel
2.3. Operational stability moves through different sectors due to the mismatch between
the reaction heat release, conduction and convective cooling
The most important aspects of the operational stability are that [26]. Localized hotspots can lead to thermal-stress-induced
(1) the OC be chemically stable and consistent as it is repeatedly fracture among the channels at the same axial location but dif-
exposed to the reducing and oxidizing environments; (2) the ferent angles in the wheel. Thick dense layers with high thermal
two-layer structure be thermally stable and resistant to thermal inertia mitigate the peripheral thermal stresses by providing a
stresses associated with temperature variations; and (3) the large heat reservoir to match the difference and thus minimize
performances of the reactor be robust to the uncertainties with the fluctuation. A faster spinning velocity and hence a shorter
the operations. cycle period also reduce the fluctuations.
332 Z. Zhao et al. / Fuel 121 (2014) 327–343

The design of the reactor should possess sufficient operational operating pressure may also affect the reactivity of the OCs, as dis-
safety margins to deal with uncertainties from the start-up of the cussed in Section 2.1.
operation, and off-design deviation during continuous operation.
The start-up period of the rotary reactor may take hours, or even 2.5. Costs of fabrication
a few days because of the large thermal inertia of the solid wheel
and the non-uniform temperature distribution along the axis of The costs of fabrication are determined mainly by the size of the
the reactor under periodic stationary states. Adequate redundancy reactor, the raw material of the OCs and the substrate, and the
should be applied to the rotary wheel and the ancillary parts (e.g., fabrication techniques. The reactor size is determined by the diam-
insulating walls, bearing) to stand the temperatures and the pres- eter, which is chosen to reach the target thermal capacity; and the
sures during the start-up period. In periodic stationary state, the height, which is selected to ensure fuel conversion. For raw OC
performances of the reactor should be robust under minor devia- materials, promising candidates are oxides of copper, nickel,
tion of the operating conditions from the set point. A sufficient cobalt, iron and manganese, because of their high fuel conversion
redundant length (e.g., 25% of the reactor height), should be and oxygen transport capabilities, among which copper, manga-
adopted near the exit of the reactor to ensure fuel conversion nese and iron are cheaper than nickel and cobalt. As for cost of
and avoid extinction of combustion. Real-time monitor of the per- the support material, alumina is among the cheapest and is easier
formances (temperature, flue species) and active controls of the to manufacture. However, materials like boron nitride, beryllium
feed flows should also be applied to effectively maintain the oxide, aluminum nitride, and silicon carbide are still the favorable
stability. choice in spite of the high cost because of their higher thermal con-
ductivity, which is crucial to the thermal stability of the reactor.
The fabrication may follow the industrial experiences of micro-
2.4. System integration channel reactors, such as the catalytic converter, the microchannel
heat exchanger [39], the microchannel Fischer–Tropsch reactor
The integration with the power generation system is the ulti- [40], etc. For example, one approach to manufacture the structural
mate goal of the design of the rotary CLC reactor. To maximize channels of the rotary wheel is the conventional powder compac-
the system efficiency, it is critical to minimize the exergy loss. tion technique. The selected ceramic powder, such as alumina,
Studies showed that for power generation systems with CLC, the mixed with appropriate binder, is spray-dried with custom dies
largest irreversibility occurs in the reactor [1]. To reduce this loss, to form the channel geometries. It is then laminated by high-pres-
it is preferred to operate the redox reactions isothermally at tem- sure hydraulic presses, and sintered at high temperature. The
peratures close to equilibrium. active OC can then be wash-coated onto the inner walls of the
For CLC reactors, the operating temperature is constrained by ceramic reactor and then calcined at high temperature to form
the material. For the OCs with low melting point (e.g., 1085 °C the porous OC layer.
for copper), supplemental firing may be required to further raise
the flue stream temperature to improve the system efficiency,
which decreases the CO2 separation efficiency. The maximum 3. Design fundamentals
allowable turbine inlet temperature (TIT) places an upper limit
on the reactor exit temperature, and supplemental cooling is To achieve the design objectives listed in Section 2, a variety of
required if the flue gas temperature is beyond the material limit parameters can be selected. The design parameters can be grouped
of the gas turbine. into three categories [22]: (1) material selection; (2) reactor de-
Isothermal operations in the fuel and air sectors are advanta- sign; (3) operating conditions. All of these parameters are highly
geous in enhancing the reaction rates, and eliminating the entropy coupled, as illustrated in Section 2, and they all strongly affect
generation associated with the heat transfer within the reactor. the performances of the reactor. Thus, it is critical to understand
Thermodynamic analysis shows that for the CLC with endothermic the fundamentals of the design, the relative importance of each
reduction, higher work output may be achieved, by operating the parameter, and the underlying logics to choose the optimal values
reduction at a lower temperature and utilizing a secondary engine of all the variables.
to extract work while transferring the heat from the air to the fuel
sector. From a practical point of view, however, the installation of 3.1. Material selection
such engine is less feasible, and hence it is preferred to operate the
redox reactions at the same temperatures (see Appendix A). 3.1.1. Porous OC layer
To reach isothermal operation in the reactor, it is critical to uti- The quality of the OC is the most important aspect for the
lize a thick dense layer with high thermal inertia and conductivity, viability of CLC. A suitable OC should possess the following charac-
that temporally stores energy from the exothermic reaction, and teristics [3,4]: (1) high reactivity; (2) reasonable stability; (3)
distribute it throughout the reactor. Preheating the feed streams abundant in nature and feasible for large-scale operations. Differ-
in the power generation system improves the isothermality by ent metal oxides have been proposed, among which, nickel
reducing the temperature difference between the gas and the solid [35,41–44], copper [41,44–48], and iron [41,49–51] are the most
near the inlet of the reactor. This can be achieved in a simple Bray- promising OC candidates. Table 2 summarizes the key properties
ton cycle with effective exhaust recuperation. The design choice and the redox reactions of these OCs. The reaction heat affects
depends on the exergy cost of non-isothermal reactor operation, the heat balance in one cycle. A highly exothermic oxidation in
the preheating options in the system, as well as the cost the air sector usually leads to a rapid rise of the reactor tempera-
implications. ture while an endothermic reduction extracts heat from the solid
The selection of the operating pressure in a CLC system depends phase to sustain the reactivity. Flow streams further absorb heat
on the power cycle. Generally, the efficiency of the system from the reactor via convective heat transfer. Thus for nickel or
increases with the compressor pressure ratio until it reaches an iron, localized hot-spots (or cold-spots) are expected in the air
optimal value, which depends on the balance between work ex- (or fuel) sector while for copper, the temperature fluctuation can
tracted from the turbine and the corresponding energy consumed be minimized by matching the redox heat release and heat convec-
in the compressor. It has been shown, at least for a combined cycle tion in both fuel and air sectors. The equilibrium conversion yield
CLC system, to be tied to the selected TIT [37,38]. Additionally, the of the OC indicates the maximum conversion of fuels. For CH4, the
Z. Zhao et al. / Fuel 121 (2014) 327–343 333

Table 2
Properties of the metal oxides used in CLC with CH4 as fuel.

CuO/Cu NiO/Ni Fe2O3/Fe3O4


Reduction CH4 + 4CuO = CO2 + 2H2O + 4Cu CH4 + 4NiO = CO2 + 2H2O + 4Ni CH4 + 12Fe2O3 = CO2 + 2H2O + 8Fe3O4
DH0 [80] 178.1 kJ/mol 156.5 kJ/mol 136.6 kJ/mol
Oxidation O2 + 2Cu = 2CuO O2 + 2Ni = 2NiO O2 + 4Fe3O4 = 6Fe2O3
DH0 [80] 312.1 kJ/mol 479.4 kJ/mol 469.4 kJ/mol
cCH4 (800/1000/1200 °C) 100%/100%/100% 100%/99%/98% 100%/100%/100%
RO 0.20 0.21 0.03
Melting point [36] 1446 °C/1085 °C 1955 °C/1455 °C 1565 °C/1597 °C
Material cost Intermediate High Low
Carbon deposition Low High Low

conversion yield is defined as cCH4 ¼ pCO2 =ðpCO2 þ pCO þ pCH4 Þ, especially for the nickel-based design, and adequate purging steam
where pi is the partial pressure of species i. All the three OCs have is necessary to clean up deposited carbon.
close to unity equilibrium conversion yields, although the yield for The porous OC layer also consists of inert materials which act as
nickel slightly decreases at 1200 °C owing to the presence of CO in a structural support maintaining the pore structures, improving
equilibrium. the gas permeability, providing a higher active surface area for
The oxygen transport capacity, defined as Roc = (mox  mred)/mox, reaction, and increasing the mechanical strength of the layer. The
where mox and mred are the masses of the fully oxidized and re- inert materials in the OC layer should possess sufficiently resis-
duced OC, respectively, describes the maximum amount of oxygen tance to thermal stresses and remain nonreactive under the redox
that can be carried by the OC between the fuel and air sectors. The environment. In addition, the inert material should also be suffi-
oxygen transport capacity decreases as inert binders are added in cient bonding yet unreactive with the active metal oxide. Potential
the OC. Roc is usually related to the cycle period, s, and the circula- options for the rotary design may include (but not limited to) BN,
tion rate of the porous OC, m _ OC , as: AlN, BeO, SiC, Al2O3, Si3N4, ZrO2, TiO2, YSZ.

_ OC ¼
Nqoc Aoc H v MO N_ f Reactivity. The redox reactivity is one of the most important prop-
m ¼ ð3Þ erties of the OC, as it directly determines the fuel conversion. Gen-
s Roc DX oc
erally, the overall redox reactions are represented by
N being the total number of channel in the wheel, doc the total thick- AðgÞ þ mr BðsÞ ! mCðsÞ þ DðgÞ ð4Þ
ness of the OC layer within each channel, qoc and Aoc the density and
the cross-sectional area of the OC layer, H the height of the reactor. where A represents fuel for reduction, and O2 for oxidation. B is the
MO is the atomic weight of oxygen, N_ f is the molar flow rate of fuel active metal oxide (or metal) within the porous layer, and mr is the
admitted into reactor, and v is the stoichiometric coefficient. DXoc is stoichiometric coefficient. As described in Section 2, the conver-
the difference in the OC conversion between the fuel and air zones. sion may invlove several key intermediate steps [4,58,59]: (1)
Copper and nickel have similar oxygen transport capacity, while the external diffusion of the reactants from the bulk of the gas
iron has a relatively lower capacity. Thus a larger OC circulation rate phase to the surface of the OCs (Fig. 4a); (2) the internal diffusion
is needed for the iron-based design to reach the same thermal of gaseous reactants through the pores of the OCs to each active
capacity. From Eq. (3) it is also obvious that for a design with a grain of the metal oxide (Fig. 4b); (3) the adsorption of the gaseous
thicker OC layer and faster rotational velocity, more fuel can be reactants onto the active metal oxide lattice surface (Fig. 5); (4)
admitted into the reactor, and hence, a smaller reactor is needed the diffusion of the reactants through the grain layer (Fig. 6); (5)
to reach a certain capacity. the intrinsic chemical reaction of the reactants. Gaseous com-
The melting point of the metal oxides is an important parame- pounds produced during the reaction go through desorption and
ter in determining the operating temperature. Because of its the diffusion via the opposite direction from steps (3)–(1).
low melting temperature, copper as the OC is usually operated un- The following discussion in this section applies simple models
der 1000 °C [4,9,10,52–54]. In addition, due to the low melting to evaluate the conversion process in which steps (1)–(5), in turn,
point, dry (or wet) impregnation method is commonly used to pre- are rate-controlling, with a focus on a plat-like porous OC
pare the copper-based OCs [4], which produces a highly porous and geometry resembling the layout of the micro-channel design.
uniform structure to provide competitive reactivities at lower tem- The characteristic time of conversion for each resistance is
perature and maintain stability during repeated cycles. However, derived and listed in Table 1. The Damköhler numbers (Da) for
the low loading content in the OC, resulting from the preparation the intermediate resistances (steps 1–4) can be calculated as
method, leads to a reduced oxygen transport capacity, Roc. the ratio of the characteristic time of that step to that of the
Carbon deposition reduces the separation efficiency, deactivates chemical reaction (step 5), and it provides an effective approach
the surface catalytic reactions by blocking the active surface areas, to identify the key parameters that ultimately determine the
and leads to operational failure, as discussed in Section 2. Experi- overall reaction rates for a specific OC. Part 2 of this series follows
ments [34,35,42,43,55,56] showed that metallic nickel, known as this method to select the design parameters and operating
catalyst to pyrolysis and Boudouard reactions, favors methane conditions suitable for three different OCs.
cracking and thus increases the tendency of carbon formation. To account for the simultaneous action of these steps, one can
On the contrary, no or very limited carbon formation has been ob- combine these resistances in series with their individual driving
served for copper- and iron-based under the operating conditions forces and fluxes, following the electrical circuit analogy, and
of interests [29,57]. In addition, carbon formation is strongly develop a detailed model to capture the spatial and temporal
dependent on the amount of oxygen available, either from the resolutions of this process. This will be addressed in depth in the
OC [22,35,36] or from steam [31,34], and becomes more prominent future.
at the end of the reduction period when the majority of the avail- Chemical kinetics control. The intrinsic rate of reaction under the
able oxygen is consumed [29,57]. Thus, sufficient oxygen should be control of chemical kinetics can be expressed as (see Appendix B for
maintained within the OC layers during the continuous operation, derivation):
334 Z. Zhao et al. / Fuel 121 (2014) 327–343

Fig. 4. Schematic illustrations of the reacting OC under the control of (a) the external diffusion, (b) the internal diffusion. The reactant is fuel (or air) for the reduction (or
oxidation).

with outer surface area of So, thickness of doc, and an initial molar
density of q ^ B (moles/m3 OC layer), the amount of metal oxide pres-
ent in the OC layer is NB ¼ ð1  X B Þq ^ B So doc , where XB is the conver-
sion of reaction. Ns is the total number of the active sites, calculated
as Ns ¼ q~ s socdocSo, with q
~ s being the concentration of the active site
on the surface and soc the specific surface area per unit volume (m2 -
grain/m3 OC layer) of the active grains. /B is the coverage of solid
reactant B in the lattice grain, and equals 1  XB. Thus, we can inte-
grate Eq. (5) to obtain the logarithmic rate law as:
q^ B
 lnð1  X B Þ ¼ t=sr ; sr ¼ ð6Þ
v r soc q~ s kr C A
From Eq. (5), a thicker OC layer (doc) and hence a larger amount of
metal oxide are recommended to improve the fuel consumption
rate (dNA/dt). As seen in Eq. (6), a higher specific surface area of
the OC (soc) and a more uniform dispersion of the metal oxide on
the porous substrate (q~ s ) should be used to reach a fast reaction rate
between the solid and the gas.
External diffusion controls. As shown in panel a of Fig. 4, if the pro-
cess is controlled by external diffusion, the concentration of reac-
tant A drops from its bulk value to zero within the gas film
adjacent to the outer surface of the porous OC layer. In this case,
the rate of OC conversion can be expressed as:
1 dN B
 ¼ So hm C A ð7Þ
mr dt
CA being the concentrations of reactant A in the bulk, and hm the
mass transfer coefficient. Thus we can obtain the linear rate law
of conversion as:
q^ B doc
X B ¼ t=sex ; sex ¼ ð8Þ
m r hm C A
For a fully developed laminar channel flow, the convective mass
transfer coefficient, hm, can be calculated as, hm = ShdDM/dh, where
Shd is the Sherwood number, and dh is the hydraulic diameter.
Fig. 5. Schematic illustrations of the adsorption and dissociation processes for (a)
NiO reduction with CH4 and (b) Ni oxidation with air.
The Sherwood number depends on the geometry of the channel.
DM is the molecular diffusivity of reactant in the bulk flow, which
1 dNB dN A can be determined using Fuller’s correlation [60]:
 ¼ ¼ kr Ns C A /B ð5Þ
mr dt dt 0:5
103 T 1:75 ðM 1 1
A þ Mi Þ
DM ¼ ð9Þ
where kr is the reaction rate coefficient, NB is the total moles of B. 1=3 1=3 2
P½ðRv ÞA þ ðRv Þi 
Here, a first-order irreversible reaction is assumed. For an OC layer
Z. Zhao et al. / Fuel 121 (2014) 327–343 335

Fig. 6. Schematic illustrations of the OC conversion under the control of the diffusion of the reactant through the grain for (a) reduction and (b) oxidation.

where T and P are the temperature and the pressure. M and Rv are From Eq. (15), it is clear that a thin OC layer significantly reduces
the molar mass and the diffusion volume, respectively. The Dam- the internal mass diffusion resistance. In addition, a higher porosity
köhler number for the external diffusion is: of the OC layer increases the effective diffusivity through the pores
and hence favors the internal diffusion.
soc q
~ s kr doc
Daex ¼ ð10Þ Adsorption process controls. As the reactant approaches the grain
hm
surface, gaseous molecules adsorb onto the lattice crystals and dis-
Eq. (10) shows that to reduce the resistance of the external mass dif- sociate while catalyze by the metal ion on the surface. In the reduc-
fusion, a small channel with a thin OC layer is recommended. In addi- tion process, the dissociated protons from the fuel bond to the
tion, a higher operating temperature with a low pressure also favors oxygen ions on the lattice surface, forming substitutional hydroxyl
the diffusivity and hence further decreases Daex. defects [63], as seen in Fig. 5a. The dissociated molecules may fur-
Internal diffusion controls. Fig. 4b shows the OC conversion process ther migrate within the lattice, break metal oxide bonds, and con-
in which the internal diffusion through the pores controls the sume the lattice oxygen ions. The metal cations may migrate
reaction rate. In this case, because of the fast external diffusion, within the grain, collapse into clusters, grow, and finally overlap
the concentration of A at the exterior surface is the same as that in to form crystallites of the solid product [58]. During the oxidation,
the bulk flow. Thus, a slowly shrinking unreacted core is expected, the oxygen molecules are ionized by the electrons within the
inside which the solid reactant B remains unreacted, as illustrated lattice, occupying active sites and forming electron holes [64], as
in Fig. 4b. The reaction occurs only on the boundary of the unreacted seen in Fig. 5b. The dissociated oxygen atom may dissolve within
core, shrinks the core size, and leaves behind a gradually thickening the grain while the metal cations may migrate to partner the
diffusion layer. Thus, the rate of conversion can be expressed on the adsorbed oxygen on the grain surface [65]. These processes are
basis of the unreacted shrinking-core model (USCM) as: complicated, and involve several intermediate steps [58,59,66].
Here, the Eley–Rideal reaction mechanism is used to qualitatively
1 dNB CA
 ¼ So D e ð11Þ model the adsorption process, as follows:
mr dt ðdoc  zun Þ
De being the effective diffusivity, and zun the thickness of unreacted AðgÞ þ SðsÞ $ ASðsÞ ð16Þ
core. By integrating Eq. (11), we can obtain the parabolic law of con-
ASðsÞ þ mr BðsÞ ! mp CðsÞ þ DðgÞ ð17Þ
version as:
where AS represents the adsorbed molecule on the lattice, and S is
q^ B d2oc
X 2B ¼ t=sin ; sin ¼ ð12Þ the open site in the lattice. Assuming that the adsorption reaction
2mr C A De
(16) controls the conversion, we can obtain (see Appendix B for
The effective diffusivity, De, can be evaluated as a combination of derivation):
the molecular and Knudsen diffusions as [61]:
  1 dNB
 ¼ kad N s C A ð18Þ
De ¼ e 2
=
1
þ
1
ð13Þ
vr dt
DK DM
where kad are the reaction rate coefficient of the forward reaction
where e is the porosity of the OC layer. The Knudsen diffusivity, DK, (16). By integrating Eq. (18), we can obtain the linear law of conver-
can be determined as [62]: sion as:
sffiffiffiffiffiffiffiffiffiffiffi
8e 2Ru T q^ B
DK ¼ ð14Þ X B ¼ t=sad ; sad ¼ ð19Þ
3soc pM A v r soc q~ s kad C A
The Damköhler number for the external diffusion can then be ex- The Damköhler number for the adsorption process can thus be ex-
pressed as: pressed as:

~ s kr d2oc
soc q kr
Dain ¼ ð15Þ Daad ¼ ð20Þ
2De kad
336 Z. Zhao et al. / Fuel 121 (2014) 327–343

As conversion proceeds, however, the available active sites, Ns The pre-exponential factor, D0, and the activation energy Ee,g, can be
changes. In addition, the nature of the sites, S, also greatly influ- obtained from experiments for each specific type of metal oxide. Eq.
ences the adsorption rate. As an example, an induction period of (22) can be rearranged and integrated to obtain the parabolic law of
NiO reduction process has been clearly observed, during which conversion as:
NiO slowly adsorbs fuel molecules and reacts to form metallic nick-
q^ B dgrain
el nuclei. As nuclei grows, an autocatalytic reduction period occurs X 2B ¼ t=sd;g ; sd;g ¼ ð24Þ
where the fuel is rapidly dissociated by the metallic Ni clusters 2mr soc C A De;g
[67,68]. Thus, during this two-stage conversion process, the active Eq. (24) is also a simplified derivation of the famous Wagner’s par-
site in the lattice alters significantly. As reduction proceeds, the abolic scaling theory [74] commonly studied in the field of metal
metallic nuclei agglomerate to form clusters, leading to the de- oxdation and corrosion.
crease of the available sites on the lattice surface. Besides, the dilu- The Damköhler number for the grain diffusion is:
ent molecules (e.g., H2O, CO2, N2) may adsorb onto the catalytic
medium, occupy the open sites and hence inhibit the adsorption q~ s kr dgrain
Dad;g ¼ ð25Þ
process [58,69]. As a result, the conversion process controlled by 2De;g
adsorption usually shows a sigmoidal conversion-versus-time pro-
Thus, thin grain layers should be used to decrease the grain diffu-
file, as documented in the literature [4,58,64]. The conversion rate
sion resistance. In addition, higher operating temperature favors
is slow initially because of the lack of catalytic nuclei, and also near
the grain diffusivity and hence increases the mass diffusion rate
the completion with few available ions present on the surface and
within the grain.
product molecules occupying the available sites. To account for
these effects, the number of the active sites can be modeled using
the Avrami–Erofeev Model (AEM) [4], as: 3.1.2. Support dense layer
The dense layer in the reactor behaves as a structure support for
ðb1Þ=b
Ns ¼ Ns;0 bð1  X B Þ½ lnð1  X B Þ ð21Þ the porous OC layer. To minimize the thermal-stress-induced fac-
tures between the porous and the dense layers, the same support
where b is the Avrami exponent indicative of the adsorption mech-
material should be used as the binder. The outer surface of the
anisms. The term, b[ln(1  XB)](b1)/b represents dynamic changes
dense layer should sustain sufficient bonding with the porous
of the active sites.
layer. The support layer should be resistant to the oxidizing and
Grain diffusion controls. As seen in Fig. 6, it is clear that the solid
corrosive environment under high temperature, and maintain ade-
reaction product (metal for reduction and metal oxidation for oxi-
quate toughness and mechanical strength to avoid fracturing dur-
dation) separates the two reactants (A and B). In order for the reac-
ing repeated cycles.
tion to further proceed, one or both reactants must penetrate the
In the periodic stationary state, the dense layer also acts as a
product layer: the dissociated ions (e.g. H+, O2) on the grain sur-
heat reservoir to match the heat transfer process: it absorbs and
face may transport into the grain lattice via defects to the unre-
temporarily stores the thermal energy released from the reactions
acted core surface, as seen in Fig. 6b [70]; the ions within the
in the porous layer, and conducts that heat axially, and finally
core (i.e., metallic cations and oxygen anions) may migrate out-
heats the bulk flow by convection. Large thermal inertia is required
wards across the product layer driven by the chemical-potential
for the dense layer to minimize the temperature fluctuation and
gradients, and enter the gas–solid surface to partner the adsorbed
avoid localized hot spots. Approximately, we can select the size
molecules and form vacancies [65].1 Besides, as the reaction pro-
of the dense layer as:
ceeds, the collapse and clustering of metallic nuclei [58] as well as
the growth stresses between the unreacted core and the product _ th
_ bulk DT max  W
cbulk m ð26Þ
layer [65] may form cracks, voids and hence porous product layers,
enabling gas phase diffusion and direct interactions between the where cbulk is the specific heat capacity of the support, W _ th is the
gas molecules and the unreacted core, as seen in Fig. 6a [71]. The dif- target thermal capacity, and DTmax is the maximum temperature
fusion mechanism by which the reactants penetrate through the fluctuation caused by the reaction heat release. m _ bulk is the
product layer is complicated and strongly depends on the type of circulation rate of the dense support solid, calculated as
the metal, the grain configuration, the impurities, and the operating m_ bulk ¼ N qbulk HAbulk =s, where qbulk and Abulk are the density and
conditions. the cross-sectional area of the bulk layer, N is the total number of
As an estimation, the unreacted shrinking-core model (USCM) is channels, and s is the cycle period. Here, the heat redistribution
widely used in the literature to simplify the grain diffusion process via conduction is neglected, and a uniform reaction heat release is
[4,58,65,68,72,73], in which a product layer is formed around an assumed within the solid phase. As an example, the design with
unreacted core inside the grain, as seen in Fig. 6. The gaseous reac- 1 MWth and Al2O3 as the dense layer requires a circulation rate of
tant is assumed to diffuse through the product layer, and react 40 kg s1 to restrict DTmax below 20 K. As predicted by Eq. (26), a
with the unreacted core. The conversion process controlled by thicker dense layer and a faster spinning velocity increase the circu-
the grain diffusion can thus be modeled as: lation rate of the solids, and hence reduce the maximum tempera-
ture fluctuation. Note that Eq. (26) tends to overestimate the
1 dNB dN A CA temperature variation by neglecting conduction, but underestimate
 ¼ ¼ soc So doc De;g ð22Þ
mr dt dt dgrain  zun;gr localized hot spots caused by non-uniform reaction heat release.
where dgrain and zun,gr are the thickness of the grain and the unre- Near the reactor inlet, the cold bulk stream is heated as it enters
acted core, respectively. The effective diffusivity can be expressed the reactor. Within this region, a rapid temperature increase is usu-
in an Arrhenius form [59,70]: ally observed in both the gas and solid phases, as seen in part 2, be-
cause of the relatively large temperature difference and hence high
De;g ¼ D0 expðEe;g =RTÞ ð23Þ heat convection rate between the solid and gas phases, as opposed
to the minimal axial temperature variation as the flow approaches
the exit. In this area the reaction heat release is relatively low be-
1
In fact, for p-type metal-deficient oxide (e.g., oxide of nickel, copper, iron), anion
cause of the low temperature, and thus, the energy in the solid
diffusion is slow, and usually cation diffusion dominates the transport process. In this phase is mainly balanced between convective cooling and conduc-
case, ‘‘A’’ in Fig. 6b should be replaced by ‘‘VB’’ (the vacancy of B) tion. Thus, a highly conductive dense layer is the key to effectively
Z. Zhao et al. / Fuel 121 (2014) 327–343 337

preheating the bulk flow. The energy balance for one cycle near the methods on the channel walls and the fabrication costs also de-
channel inlet can be expressed as: pend on the geometry.
  The sector design should provide adequate residence time for
d dT s the channel in each sector. The channel residence time is deter-
Abulk ks ’ Pc hgs ðT s  T g Þ ð27Þ
dz dz mined by the rotational velocity (x), and the size of each sector
(hi), i.e., si = hi/x. For the optimal design, a large fuel sector is al-
where Abulk, Pc are the cross-sectional area of the solid phase, and ways preferred as more fuel can be admitted into the reactor and
the channel perimeter. Ts, and Tg are the temperature of the solid thus a smaller reactor is needed to reach the thermal capacity
and gas phases. ks, and hgs are the thermal conductivity of the sup- (Eq. (32)). However, complete reduction of the OC should be
port substrate and the convective heat-transfer coefficient. z is the avoided in selecting the fuel sector size, since the reduction rate
channel axial direction. Eq. (27) indicates that a thicker dense layer may slow down as the OC is consumed and the fuel conversion
(Abulk) with a higher thermal conductivity (ks) sustains stronger con- in the channel drops significantly as the OC approaches full reduc-
vective cooling near the inlet, reduces the axial temperature varia- tion. The residual metal oxide in the OC also controls carbon depo-
tion of the solid, favors the preheating process, and thus enhances sition risk in the fuel sector. Thus, the size of the fuel sector should
the solid temperature as well as the reaction rates. Some ceramic satisfy:
materials with good potential because of the high thermal conduc-
tivity are listed in Table 3. Note that the thermal conductivity may
sfuel ¼ hfuel =x 6 sfull-red ð28Þ
vary significantly with respect to the temperature, crystalline struc- On the other hand, the size of the air sector should be large enough
tures of the material, and the preparation method. such that the OC on the walls are sufficiently exposed to and
re-oxidized by the flowing air. Complete oxidation of the OC may
not be necessary due to the fact that the oxidation rate may de-
3.2. Design parameters
crease significantly as the OC approaches complete oxidation state,
as noted in Part 2. In addition, the catalytic effects of the metal in
The design of the channel determines the flow and the reaction
the OC layer may help enhance the adsorption and dissociation of
patterns within the reactor. A small channel (dh) with a high spe-
the fuel molecules on the grain lattice and hence improve the
cific surface area (sc) is required for the channel to provide suffi-
reduction rate in the following fuel sector, as discussed in
cient contact between the flow and the OC, and hence, enhance
Section 3.1. Thus, it is reasonable to maintain a certain amount of
the reactivities. The size of the channel (dh) affects the mass trans-
OC unconverted in the OC layer at the end of the air sector. The size
fer rate (hm), as discussed in Section 3.1. On the other hand, a small
of the air sector should be selected such that:
channel also improves the convective cooling, i.e., hgs = Nu  kg/dh,
where Nu is the Nussult number, leading to a lowered solid tem- sair ¼ hair =x P ssufficient-ox ð29Þ
perature and then reduced redox rates, as well as an enlarged local
temperature fluctuation caused by the energy mismatch, and a lar-
sfull-red, and ssufficient-ox are mainly controlled by the redox character-
istics of the OC under specific operating conditions.
ger axial temperature difference near the inlet (see Eq.(27)). In this
To ensure separation of CO2 and avoid dilution of air, the resi-
case, a much thicker solid layer is required to provide sufficient
dence time of channels (si) in each purging sector should be kept
heat for convection in the preheat zone and maintain the energy
adequately longer than that for steam (ti) to purge through the
balance within the reactor. Because of these two aspects, a reason-
channel:
able channel size (dh) for the rotary CLC reactor ranges from 0.5 to
5 mm. Thus, the flow field within the channel is in general laminar. si ¼ hi =x > ti ; i ¼ fuel purge or air purge ð30Þ
The pressure drop within the reactor is caused by skin friction,
2 where ti is determined by Eq. (2). Therefore, the choice of the purge
and can be calculated as [75], Dp ¼ 12 C f ug lg H=dh , where ug, and lg
sector sizes (hi) is usually coupled with the selection of the purge
are the mean velocity and the dynamic viscosity of the bulk flow, Cf
velocity and the channel height. As discussed in Section 2.2, one
is the friction constant from the friction factor (f = Cf/Re). For a
can reduce the purging volume (hi) by increasing the steam veloci-
squared-channel channel with the width of 2 mm and the height
ties while maintaining a constant rotational speed. This is usually
of 1 m, a 1 kPa pressure drop is expected for the operation at
the case for the air purging sector where we minimize the purging
1000 °C and ug of 1 m/s. In general, the pressure drop is negligible
volume by admitting a fast stream. However, the steam velocity in
as compared to the operating pressure.
the fuel purge sector is usually constrained in order to oxidize all
The channels can be formed in a variety of geometries, e.g., grid
the residual fuel within the channel. Thus, in the case with low
type, honeycomb geometry shapes, plate types, etc. Different
reduction reactivity, the purge volume can be much larger (see
geometries of the channel are accompanied by different Sh’s,
the iron-based case in Part 2).
Nu’s, and Cf’s and hence affect the mass and heat transfer processes
For the rotary design with different OCs, the rate-limiting step
as well as the pressure distribution. Besides, the geometry may
varies, and thus the size requirements of each sector differ accord-
affect the mechanical strength and alter the resistance to the
ingly. For instance, as showed in Part 2, the nickel-based design has
stresses associated with temperature variations. The OC coating
a much larger air sector owing to its slow oxidation rate; the longer
gas residence time needed for fuel conversion requires a much lar-
Table 3 ger share of fuel purge sector in the iron-based design for separa-
Thermal conductivity of ceramic materials suitable for the rotary CLC reactor. tion; for copper-based design, the size of the fuel sector takes the
Conductivity (W m1 K1) Refs. majority of the reactor and it is constrained by the full reduction
of the OC.
BN Boron nitride 200–600 [81]
BeO Beryllium oxide 50–200 [82] The height of the reactor is selected to ensure fuel conversion:
AlN Aluminum nitride 50–200 [83]
H P ð1 þ jÞz99% ð31Þ
SiC Silicon carbide 50–300 [84]
Al2O3 Aluminum oxide 10–30 [82] where z99% is the height required to reach 99% fuel conversion, and
Si3N4 Silicon nitride 10–30 [85]
j is the redundancy factor. Generally, a 25% redundancy is required
Note: Typical values under the conditions of interest for rotary CLC. The data varies to account for the uncertainties with the operations. The diameter
with temperature, material structure, and preparation method. of the reactor is determined to reach the target thermal capacity:
338 Z. Zhao et al. / Fuel 121 (2014) 327–343

1 longer residence time of the gas is needed to consume the residual


hfuel D2 n_ f Dhr ¼ W
_ th ð32Þ
8 fuel in the channel. Thus, the height of the reactor is reduced at the
cost of an enlarged reactor diameter. The optimal choice depends
Here D is diameter of the rotary wheel, Dhr is the enthalpy of reac-
on the tradeoff effect and it differs with the selection of the OC.
tion of the fuel, and n_ f is the fuel flow rate in the feed chamber. Eq.
The rotational velocity (x) controls the cycle period (s). A faster
(32) shows that the thermal capacity of the reactor scales with the
rotational speed is preferred to increase the admitted fuel flow rate
square of the diameter. The height of the reactor remains un-
(Eq. (3)), reduce the temperature fluctuation (Eq. (26)), decrease
changed with the up-scaling.
the carbon deposition risk and maintain the fast reduction reactiv-
ity (Eq. (28)). On the other hand, sufficient residence time is
3.3. Operating condition required for complete residual gas cleaning (Eq. (30)) and the OC
regeneration (Eq. (29)). Therefore, the choice of the rotational
The operating temperature is one of the most important factors velocity is to minimize the cycle period while maintain long
for the overall performances of the rotary reactor. The inlet (Tin) enough residence time for each channel in the purge sectors and
and the exit (Tex) temperatures of the rotary reactor are selected the air sector.
on the basis of the characteristics of the OCs and the compatibility The feed velocity (ufuel) in the fuel sector determines the resi-
requirements from the power generation system. As discussed pre- dence time of gas and hence the required height (z99%) for complete
viously, a higher operating temperature always enhances the redox fuel conversion (Eq. (1)). A higher feed velocity reduces the diam-
rates, improves the fuel conversion, and reduces the size of the eter of the reactor by increasing the feed rate of the fuel (Eq. (32))
reactor and the costs of fabrication. From the system perspective, but requires a longer reactor height (z99%) to convert the fuel. The
a higher temperature corresponds to a higher turbine inlet temper- selection of the fuel velocity is usually related to that of the fuel
ature (TIT), which leads to a higher net plant efficiency. However, concentration at the inlet, as discussed previously, and these two
the inlet temperature of the bulk flow is constrained by the capa- parameters should be determined together to reach high fuel feed
bility of the preheating units in the system, as discussed in Sec- rate and maintain a reasonable reactor height for fuel conversion.
tion 2.4. The selection of the exit temperature from the reactor is The fuel purge velocity (ufuel,p) is then selected accordingly to pro-
mainly controlled by the material limits from the OCs and the vide a similar flow field within the channel to completely convert
gas turbine. For the state-of-the-art gas turbines used in the con- the residual fuel. As a result, for the design with a slow fuel veloc-
ventional power plants, the TIT can be 1400 °C or above [76]. It is ity, a low purge rate should be used in the following purge sector,
worth noting that for fluidized CLC reactor with OCs in the form leading to a larger share of the fuel purge sector required for sep-
of particles, the operating temperatures are usually constrained aration (Eq. (30)). In contrast, the steam velocity in the air purge
below 1200 °C [4], because of the concerns with agglomeration sector (uair,p) can be much higher, as pointed out in Section 2.2,
and defluidization. In contrast, these concerns are less significant and thus, the air purge sector can be much smaller.
for the rotary reactor because of the regulated flow patterns The air feed velocity (uair) is mainly determined by the overall
through the reactor and the limited temperature variations in the heat balance:
cyclic operations. Thus, a reasonable operating temperature for ro-
tary CLC reactor is in the range of 1000–1400 °C. Similar inlet tem- hair quair cp ðT ex  T in Þ  hfuel n_ f Dhr ð33Þ
peratures are recommended for the feed streams to reduce the
where q and cp are the density and the specific heat capacity of the
thermal stresses in the reactor, which on the other hand may lead
air stream. Sufficient air stream is required to transfer the reaction
to a more complex system design and higher operational cost.
heat and maintain the operating temperature. The air feed velocity
The operating pressure (P) is also critical in determining the
also affects the oxygen concentration within the reactor and thus
functionality of the reactor. In general, the pressure is determined
affects the characteristic time for re-oxidation (ssufficient-ox). Thus, uair
by the requirements from the power generation system, with the
and hair should be determined simultaneously on the basis of Eqs.
consideration of the net plant efficiency, the system specific work
(29) and (33).
output and the CO2 compression. A reasonable operating pressure
ranges from 10–40 atm. The pressure affects the reaction rates by
changing the gas concentration, the diffusion and adsorption pro- 4. Design procedure
cesses. The reaction rate may increase with a higher total pressure
because of the increase in reactant concentration, as reported by The design procedure (Fig. 7) is based on the relative impor-
Ref. [77–79]. On the other hand, the decrease in the molecular dif- tance and the dependent relationships of the design parameters,
fusivity at the high pressure (Eq. (9)) may increase the external and as discussed in Section 3.
internal mass diffusion resistances and thus inhibit the reaction As showed in Fig. 7, the design input includes the target thermal
rates. The amount of the available open sites on the grain lattice capacity (W _ th ), the type of the fuel, the operating temperatures
may decrease at elevated pressures, as discussed in Appendix B, (Tin, Tex) and the pressure (P), and the type of the metal oxide.
which leads to a lower reaction rate. Besides, the operating pres- The design input data are determined by the compatibility require-
sure also influences the heat convection by changing the density ments from the system integration. The chemical and thermal
and hence the thermal inertia of the streams. properties of the metal oxide must allow continuous cycling in
The feed stream in the fuel sector is usually diluted by recircu- the reactor at the operating temperature and pressure without
lated CO2 to control the fuel feed rate in the channel and lower the melting, fouling, and decomposition.
operating temperature. The fuel concentration (xfuel,0) at the inlet The OC design and the preparation techniques are then deter-
affects the reduction rate, as expressed in Eq. (6), and changes mined to provide sufficient reduction and oxidation rates. A thick
the reactor height required to convert the fuel (Eq. (31)). One can OC layer (doc) with large active surface areas (soc) leads to a higher
always reduce the reactor height by using a higher fuel concentra- consumption rate of the fuel (Eqs. (3) and (5)). On the other hand,
tion and a low flow velocity, as pointed out in Section 2.1. How- doc is constrained to minimize the mass transfer resistance caused
ever, the fuel stream with high concentration near the inlet by the external and internal diffusion within the porous layer (Eqs.
rapidly consumes the available OCs, leading to a shortened reduc- (10) and (15)). A highly porous (e) OC layer improves the gas per-
tion period (Eq. (28)), and hence an enlarged reactor size (Eq. (32)). meability through the pores (Eq. (13)) and thin grain layers (dgrain)
A large fuel purge sector is also necessary in this case because a on the porous substrate enhances the diffusion through the
Z. Zhao et al. / Fuel 121 (2014) 327–343 339

Fig. 7. Design procedure of the rotary CLC reactor.

product layer in the lattice (Eq. (25)). An intermediate range of the (Eq.(28)). The choices for the size (sair) and the feed flow rate (uair)
OC conversion (DXoc) is preferred to maintain fast redox rates of the air sector are governed by the overall energy balance of the
during the cyclic operation (Eq. (21)) and eliminate the carbon reactor (Eq. (33)) and the residence time required to reach suffi-
deposition risk, while a sufficient oxygen transport capacity (Roc) cient OC regeneration (Eq. (29)). Depending on the reactor height
is required for full fuel consumption (Eq. (3)). The selection of and the steam velocity, the necessary size of the purge sectors
the support material and the design of the dense layer should pro- (sfuel,p or sair,p) can then be selected (Eq. (30)). The rotational veloc-
vide sufficient thermal inertia and conductivity to minimize the ity (x) and the share of each sector (hi) are then selected according
temperature fluctuation (Eq. (26)) and maintain reasonable iso- to sfuel, sair, sfuel,p and sair,p. A higher steam rate (uair,p) is recom-
thermality of the reactor (Eq. (27)). The support material should mended in the air purge sector to minimize hair,p. Finally, the diam-
also be stable and resistant to corrosion and thermal distortion. eter of the reactor can be determined on the basis of the fuel feed
The choice for the channel design is based on the redox reactiv- rate and the target thermal capacity (Eq. (32)).
ities of the OC, the flow pattern required for the reaction, the
dimensions of the OC layer and the dense support layer. A suffi- 5. Conclusion
ciently large channel should be utilized to accommodate both lay-
ers, maintain high temperature near the inlet (Eq. (27)), and reduce In this work, the design fundamentals of a novel rotary CLC
the pressure drop along the reactor. On the other hand, the channel reactor with micro-channel structures are examined in details.
width is constrained below a certain value to reduce the external The primary design objectives and criteria of the rotary reactor
mass diffusion resistance (Eq. (10)). The geometry of the channel have been proposed, including the requirements on fuel conver-
affects the mechanical stability as well as the fabrication costs of sion, CO2 separation, operational stabilities, the integration and
the reactor. the economic feasibility in a power generation system. For each
The operating conditions in the fuel and the fuel purge sectors of the objective, the key parameters that control the performance
are then selected on the basis of the reduction kinetics of the OC. are identified.
The reactor height (z99%) for fuel conversion is an important indica- To achieve the design objectives, a variety of parameters can
tor for the choice of the feed fuel concentration (xfuel,0) and the flow be determined, including material selection, reactor geometry
velocity (ufuel). With the same amount of fuel admitted, z99% de- and arrangement, and the operating conditions. All of these
creases with the increase of the fuel concentration at the inlet, parameters are highly coupled, and strongly affect the perfor-
which on the other hand may lead to a shorter reduction time in mance of the reactor. A thorough discussion of the design funda-
the fuel sector before complete reduction occurs. A similar velocity mentals is presented in this study, which provides insights to the
(ufuel,p) should be used in the fuel purge sector to completely design optimization:
oxidize the residual fuel. The reactor height can then be selected
based on z99% with sufficient redundancy (Eq. (31)). – The quality of the OC is the most important design factor. A
The channel residence time in the fuel sector (sfuel) is suitable OC exhibits high redox rates and maintains stability
constrained by the characteristic time for complete reduction in repeated cycles. The redox of the OC involves multiple
340 Z. Zhao et al. / Fuel 121 (2014) 327–343

intermediate steps, and the key parameters are identified. A methodology to the design of three commonly used OCs, i.e., nick-
thick dense support layer with high thermal inertia and conduc- el, copper, and iron, and evaluates the behaviors of the rotary reac-
tivity is necessary to maintain the stability of the reactor. tor using numerical simulations.
– The design of the channel influences the heat and the mass
transfer between the bulk flow and the OC. The sizes of the dif- Acknowledgements
ferent sectors are governed by the characteristic time for reduc-
tion, oxidation, and purging. The dimensions of the reactor are This study is financially supported by a grant from the MASDAR
controlled by fuel conversion and the thermal output. Institute of Science and Technology and the King Abdullah Univer-
– The pressure and temperature are the most important sity of Science and Technology (KAUST) Investigator Award.
operating parameters and they are determined by the compat-
ibility requirement from the system integration. The inlet
concentration and the flow rate in fuel sector directly control Appendix A.
the height of the reactor while the air flow rate in the air
sector determines the overall heat balance. The feed velocities The exergy or availability of a system measures the theoretical
in purge sectors are determined to ensure complete CO2 maximum work output during a thermal cyclic process. In a typical
separation. CLC setup, the oxidation reaction is highly exothermic (DHox < 0)
while the reduction reaction is endothermic (DHred > 0) or mildly
The design procedure has been proposed on the basis of the exothermic (DHred < 0). Additionally, the two reactors may be oper-
influences, the relative importance, and the dependent relation- ated under different temperatures, and hence it is theoretically
ships of the design parameters with respect to the performance possible to install a secondary Carnot engine between the two
of the rotary reactor. Part 2 presents the application of this reactors that extracts additional work while reversibly transferring
heat from the oxidation to the reduction reactor.

Fig. A1. Schematic drawings of (a) an ideal CLC reactor system (b) the entropy generation in practical CLC systems.
Z. Zhao et al. / Fuel 121 (2014) 327–343 341

In panel a of Fig. A1, we assume that the two reactors operate If we assume that the adsorbed surface species, AS, and S are in
reversibly and isothermally and thus can be treated as heat reser- quasi-steady states, therefore,
voirs. Exhaust streams from the reactors are used to preheat the
k1 C A hS
feed fuel and air streams in ideal heat exchangers. The heat hAS ¼ ðB1Þ
exchangers are also assumed to be balanced so that the inlet
k1 þ k2 /B
streams leaving the exchanger are perfectly heated up to the where k1 and k1 are the forward and reverse rate coefficients of
respective reactor temperatures. The heat extracted/added to reaction (16), and k2 is the forward reaction rate of reaction (17).
maintain an isothermal reaction is equal to the heat of that reac- hS, and hAS are the coverage ratios of the available active sites, and
tion. Hence, DHox = QH and DHred = QC. Therefore, DHr = - the sites occupied by AS. The summation of hS and hAS equals unity.
(DHox + DHred) = QH + QC. The temperatures of the air reactor Thus,
and the fuel reactor are TH, and TC, respectively. Applying the first
and second laws of thermodynamics to the control volume (inside k1 þ k2 /B
hS ¼ ðB2Þ
the dashed box), the maximum work for cases where the reduction k1 þ k2 /B þ k1 C A
reaction is endothermic or exothermic is given by
Thus, the overall reaction rate is:
   
T0 1 1 1 dNB Ns k1 k2 C A /B
W max;ex ¼ jDHr j 1   jDHred jT 0  ðA1Þ  ¼ ðB3Þ
TH TC TH vr dt k1 þ k2 /B þ k1 C A
   
T0 1 1 By assuming the relative importance of the adsorption and the
W max;en ¼ jDHr j 1  þ jDHred jT 0  ðA2Þ chemical reaction steps, we can simplify Eq. (B3) into:
TH TC TH
From Eq. (A1), it is obvious that the work output for the CLC system 1 dNB
 ¼ k1 Ns C A ; adsorption controls ðB4Þ
with exothermic reduction reaction is highest when TH = TC. In this vr dt
case, both reactors supply the energy for the power generation, and
hence a higher reduction temperature (TC) provides heat with a 1 dNB Ns K 1 k2 C A /B
 ¼ ; chemical reaction controls ðB5Þ
higher quality. vr dt 1 þ K 1CA
For the case with endothermic reduction reaction, Eq. (A2) sug-
where K1 is the equilibrium constant of reaction (16). If K1CA  1,
gests that the system work output increases with the difference
Eq. (B5) becomes identical to Eq. (5). In this case, the reaction rate
between the oxidation and the reduction reactor temperatures.
rises with the increase of the reactant concentration, and the con-
This is because the two steps in CLC upgrade the low quality heat
version is mainly controlled by the amount of available reactant
(at TC) from the fuel reactor to the air reactor (at TH) and more high-
in the bulk phase. On the other hand, if the reaction proceeds with
quality energy can be extracted in the air reactor at a higher tem-
a high concentration of the reactant, i.e., K1CA 1, thus Eq. (B5)
perature [1,86,87]. Therefore, theoretically, extra work output can
becomes:
be produced via the secondary engine as the heat is transferred
from the air reactor (hot reservoir) to the fuel reactor (cold reser- 1 dNB
 ¼ Ns k2 /B ðB6Þ
voir) to maintain the overall energy balance (see panel a of Fig. A1). vr dt
However in a practical CLC system, it is less likely to install an
In this case, the active sites on the grain lattice are saturated by the
engine between the two reactors to exchange heat while extracting
reactant, and a further increase in the reactant concentration has
work because the transferred energy is mainly in the form of the
little effect on the reaction rate. Here the reaction is mainly con-
internal energy of the solids and it is transferred via the physical
trolled by the available sites (hS) on the surface. In addition, the
transport of the solids between the two reactors. Thus, in the
product molecules in the reduction reaction, e.g., H2O, CO2, may ad-
absence of a secondary engine, the direct heat transfer results in
sorb onto the grain, block the active sites on the lattice surface, and
exergy destruction proportional to the temperature difference.
hence further reduce hS. These effects may prevail at elevated
The entropy generation can be estimated as completely due to
pressures.
the heat transfer between the reactors and is given by
 
1 1 References
S_ gen ¼ DHred  ðA3Þ
TC TH
[1] Richter HJ, Knoche KF. Reversibility of combustion processes. In: ACS
Thus, the maximum work output in this case (panel b of Fig. A1) is symposium series; 1983. p. 71–85.
  [2] Lyngfelt A, Leckner B, Mattisson T. A fluidized-bed combustion process with
T0 inherent CO2 separation; application of chemical-looping combustion. Chem
W max;en ¼ jDHr j 1  ðA4Þ Eng Sci 2001;56:3101–13.
TH [3] Hossain MM, de Lasa HI. Chemical-looping combustion (CLC) for inherent CO2
separations – a review. Chem Eng Sci 2008;63:4433–51.
Therefore, in this case, the maximum work output is independent [4] Adanez J, Abad A, Garcia-Labiano F, Gayan P, de Diego LF. Progress in chemical-
on the selection of the temperature difference, and the thermody- looping combustion and reforming technologies. Prog Energy Combust Sci
namic advantage of operating the fuel reactor at a lower tempera- 2012;38:215–82.
[5] Johansson M, Mattisson T, Lyngfelt A. Use of NiO/NiAl2O4 particles in a 10 kW
ture disappears owing to the irreversibility associated with the chemical-looping combustor. Ind Eng Chem Res 2006;45:5911–9.
heat transfer. From a practical point of view, it is preferred to oper- [6] Linderholm C, Abad A, Mattisson T, Lyngfelt A. 160 h of chemical-looping
ate the fuel reactor at the same temperature to reach a high combustion in a 10 kW reactor system with a NiO-based oxygen carrier. Int J
Greenhouse Gas Control 2008;2:520–30.
reduction rate in the fuel reactor, and to minimize the thermal
[7] Abad A, Mattisson T, Lyngfelt A, Ryden M. Chemical-looping combustion in a
stresses caused by the temperature variations during the transport 300W continuously operating reactor system using a manganese-based
of the solids. oxygen carrier. Fuel 2006;85:1174–85.
[8] Berguerand N, Lyngfelt A. Design and operation of a 10 kWth chemical-looping
combustor for solid fuels – testing with South African coal. Fuel
Appendix B. 2008;87:2713–26.
[9] Adanez J, Gayan P, Celaya J, de Diego LF, Garcia-Labiano F, Abad A. Chemical
looping combustion in a 10 kWth prototype using a CuO/Al2O3 oxygen carrier:
The adsoprtion, and the chemical reaction processes can be effect of operating conditions on methane combustion. Ind Eng Chem Res
described by Eley–Rideal reaction mechanism (Eqs. (16) and (17)). 2006;45:6075–80.
342 Z. Zhao et al. / Fuel 121 (2014) 327–343

[10] de Diego LF, Garcia-Labiano F, Gayan P, Celaya J, Palacios JM, Adanez J. [41] Adanez J, de Diego LF, Garcia-Labiano F, Gayan P, Abad A, Palacios JM. Selection of
Operation of a 10 kWth chemical-looping combustor during 200 h with a oxygen carriers for chemical-looping combustion. Energy Fuels 2004;18:371–7.
CuO–Al2O3 oxygen carrier. Fuel 2007;86:1036–45. [42] Johansson M, Mattisson T, Lyngfelt A, Abad A. Using continuous and pulse
[11] Dueso C, Garcia-Labiano F, Adanez J, de Diego LF, Gayan P, Abad A. Syngas experiments to compare two promising nickel-based oxygen carriers for use in
combustion in a chemical-looping combustion system using an impregnated chemical-looping technologies. Fuel 2008;87:988–1001.
Ni-based oxygen carrier. Fuel 2009;88:2357–64. [43] Gayan P, de Diego LF, Garcia-Labiano F, Adanez J, Abad A, Dueso C. Effect of
[12] Kolbitsch P, Bolhar-Nordenkampf J, Proell T, Hofbauer H. Operating experience support on reactivity and selectivity of Ni-based oxygen carriers for chemical-
with chemical looping combustion in a 120 kW dual circulating fluidized bed looping combustion. Fuel 2008;87:2641–50.
(DCFB) unit. Int J Greenhouse Gas Control 2010;4:180–5. [44] Cho P, Mattisson T, Lyngfelt A. Comparison of iron-, nickel-, copper- and
[13] Kolbitsch P, Proell T, Bolhar-Nordenkampf J, Hofbauer H. Design of a chemical manganese-based oxygen carriers for chemical-looping combustion. Fuel
looping combustor using a dual circulating fluidized bed reactor system. Chem 2004;83:1215–25.
Eng Technol 2009;32:398–403. [45] de Diego LF, Garcia-Labiano F, Adanez J, Gayan P, Abad A, Corbella BM, et al.
[14] Shen L, Wu J, Gao Z, Xiao J. Reactivity deterioration of NiO/Al2O3 oxygen carrier Development of Cu-based oxygen carriers for chemical-looping combustion.
for chemical looping combustion of coal in a 10 kWth reactor. Combust Flame Fuel 2004;83:1749–57.
2009;156:1377–85. [46] Corbella BM, de Diego L, Garcia F, Adanez J, Palacios JM. The performance in a
[15] Gu H, Shen L, Xiao J, Zhang S, Song T. Chemical looping combustion of biomass/ fixed bed reactor of copper-based oxides on titania as oxygen carriers for
coal with natural iron ore as oxygen carrier in a continuous reactor. Energy chemical looping combustion of methane. Energy Fuels 2005;19:433–41.
Fuels 2011;25:446–55. [47] Garcia-Labiano F, Adanez J, de Diego LF, Gayan P, Abad A. Effect of pressure on
[16] Li F, Zeng L, Velazquez-Vargas LG, Yoscovits Z, Fan LS. Syngas chemical looping the behavior of copper-, iron-, and nickel-based oxygen carriers for chemical-
gasification process: bench-scale studies and reactor simulations. AIChE J looping combustion. Energy Fuels 2006;20:26–33.
2010;56:2186–99. [48] Chuang SY, Dennis JS, Hayhurst AN, Scott SA. Development and performance of
[17] Fan LS. Chemical looping systems for fossil energy conversions. Hoboken, New Cu-based oxygen carriers for chemical-looping combustion. Combust Flame
Jersey: John Wiley & Sons Inc.; 2010. 2008;154:109–21.
[18] Noorman S, van Sint Annaland M, Kuipers H. Packed bed reactor technology for [49] Johansson M, Mattisson T, Lyngfelt A. Investigation of Fe2O3 with MgAl2O4 for
chemical-looping combustion. Ind Eng Chem Res 2007;46:4212–20. chemical-looping combustion. Ind Eng Chem Res 2004;43:6978–87.
[19] Noorman S, Gallucci F, van Sint Annaland M, Kuipers JAM. Experimental [50] Mattisson T, Johansson M, Lyngfelt A. Multicycle reduction and oxidation of
investigation of chemical-looping combustion in packed beds: a parametric different types of iron oxide particles-application to chemical-looping
study. Ind Eng Chem Res 2011;50:1968–80. combustion. Energy Fuels 2004;18:628–37.
[20] Dahl IM, Bakken E, Larring Y, Spjelkavik AI, Håkonsen SF, Blom R. On the [51] Ryden M, Cleverstam E, Johansson M, Lyngfelt A, Mattisson T. Fe2O3 on Ce-, Ca-,
development of novel reactor concepts for chemical looping combustion. or Mg-stabilized ZrO2 as oxygen carrier for chemical-looping combustion using
Energy Proc 2009;1:1513–9. NiO as additive. AIChE J 2010;56:2211–20.
[21] Håkonsen SF, Blom R. Chemical looping combustion in a rotating bed reactor – [52] Garcia-Labiano F, de Diego LF, Adanez J, Abad A, Gayan P. Reduction and
finding optimal process conditions for prototype reactor. Environ Sci Technol oxidation kinetics of a copper-based oxygen carrier prepared by impregnation
2011;45:9619–26. for chemical-looping combustion. Ind Eng Chem Res 2004;43:8168–77.
[22] Zhao Z, Chen T, Ghoniem AF. Rotary bed reactor for chemical-looping [53] Forero CR, Gayán P, García-Labiano F, de Diego LF, Abad A, Adánez J. High
combustion with carbon capture. Part 1: reactor design and model temperature behaviour of a CuO/cAl2O3 oxygen carrier for chemical-looping
development. Energy Fuels 2013;27:327–43. combustion. Int J Greenhouse Gas Control 2011;5:659–67.
[23] Zhao Z. Rotary bed reactor for chemical-looping combustion with carbon [54] Gayán P, Forero CR, Abad A, de Diego LF, García-Labiano F, Adánez J. Effect of
capture. Dept. of Mechanical Engineering, Massachusetts Institute of support on the behavior of Cu-based oxygen carriers during long-term CLC
Technology, Cambridge; 2012. operation at temperatures above 1073 K. Energy Fuels 2011;25:1316–26.
[24] Pavone D. CO2 capture by means of chemical looping combustion. In: [55] Jerndal E, Mattisson T, Lyngfelt A. Investigation of different NiO/NiAl2O4
Proceedings of the COMSOL multiphysics user’s conference, Paris, France; 2005. particles as oxygen carriers for chemical-looping combustion. Energy Fuels
[25] Pavone D, Rolland M, Lebas E. CO2 capture using chemical looping combustion 2009;23:665–76.
for gas turbine application. In: Proceedings of 8th international conference of [56] Adanez J, Garcia-Labiano F, de Diego LF, Gayan P, Celaya J, Abad A. Nickel-
greenhouse gas control technologies (GHGT-8), Trondheim, Norway; 2006. copper oxygen carriers to reach zero CO and H2 emissions in chemical-looping
[26] Zhao Z, Chen T, Ghoniem AF. Rotary bed reactor for chemical-looping combustion. Ind Eng Chem Res 2006;45:2617–25.
combustion with carbon capture. Part 2: base case and sensitivity analysis. [57] de Diego LF, Gayan P, Garcia-Labiano F, Celaya J, Abad A, Adanez J.
Energy Fuels 2013;27:344–59. Impregnated CuO/Al2O3 oxygen carriers for chemical-looping combustion:
[27] Zhao Z, Ghoniem AF. Design of a rotary reactor for chemical-looping avoiding fluidized bed agglomeration. Energy Fuels 2005;19:1850–6.
combustion. Part 2: Comparison of copper-, nickel-, and iron-based oxygen [58] Richardson JT, Scates RM, Twigg MV. X-ray diffraction study of the hydrogen
carriers. Fuel 2014;121:344–60. reduction of NiO/a-Al2O3 steam reforming catalysts. Appl Catal A 2004;267:35–46.
[28] Aslan-zada FE, Mammadov VA, Dohnal F. Brush seals and labyrinth seals in gas [59] Ward TL, Dao T. Model of hydrogen permeation behavior in palladium
turbine applications. Proc Inst Mech Engrs, Part A: J Power Energy membranes. J Membr Sci 1999;153:211–31.
2013;227:216–30. [60] Fuller EN, Schettler PD, Giddings JC. New method for prediction of binary gas-
[29] Cho P, Mattisson T, Lyngfelt A. Carbon formation on nickel and iron oxide- phase diffusion coefficients. Ind Eng Chem 1966;58:18–27.
containing oxygen carriers for chemical-looping combustion. Ind Eng Chem [61] Wakao N, Smith JM. Diffusion in catalyst pellets. Chem Eng Sci
Res 2005;44:668–76. 1962;17:825–34.
[30] Jin H, Okamoto T, Ishida M. Development of a novel chemical-looping [62] Peterson EE. Chemical reaction analysis. Engelwood Cliffs, New
combustion: synthesis of a solid looping material of NiO/NiAl2O4. Ind Eng Jersey: Prentice-Hall; 1965.
Chem Res 1999;38:126–32. [63] Liu Y, Tan X, Li K. Mixed conducting ceramics for catalytic membrane
[31] Ishida M, Jin H, Okamoto T. Kinetic behavior of solid particle in chemical- processing. Catal Rev 2006;48:145–98.
looping combustion: suppressing carbon deposition in reduction. Energy Fuels [64] Hossain MM, de Lasa HI. Reactivity and stability of Co–Ni/Al2O3 oxygen carrier
1998;12:223–9. in multicycle CLC. AIChE J 2007;53:1817–29.
[32] Jin H, Ishida M. Reactivity study on natural-gas-fueled chemical-looping [65] Birks N, Meier GH, Pettit FS. Introduction to the high temperature oxidation of
combustion by a fixed-bed reactor. Ind Eng Chem Res 2002;41:4004–7. metals. Cambridge University Press; 2006.
[33] Hoteit A, Chandel MK, Delebarre A. Nickel- and copper-based oxygen carriers [66] Hossain MM, de Lasa HI. Reduction and oxidation kinetics of Co–Ni/Al2O3
for chemical looping combustion. Chem Eng Technol 2009;32:443–9. oxygen carrier involved in a chemical-looping combustion cycles. Chem Eng
[34] Chandel MK, Hoteit A, Delebarre A. Experimental investigation of some metal Sci 2010;65:98–106.
oxides for chemical looping combustion in a fluidized bed reactor. Fuel [67] Bandrowski J, Bickling CR, Yang KH, Hougen OA. Kinetics of the reduction of
2009;88:898–908. nickel oxide by hydrogen. Chem Eng Sci 1962;17:379–90.
[35] Mattisson T, Johansson M, Lyngfelt A. The use of NiO as an oxygen carrier in [68] Szekely J, Evans J, Sohn H. Gas-solid reactions. New York, NY: Academic Press;
chemical-looping combustion. Fuel 2006;85:736–47. 1976.
[36] Jerndal E, Mattisson T, Lyngfelt A. Thermal analysis of chemical-looping [69] Richardson JT, Lei M, Turk B, Forster K, Twigg MV. Reduction of model steam
combustion. Chem Eng Res Des 2006;84:795–806. reforming catalysts: NiO/a-Al2O3. Appl Catal A 1994;110:217–37.
[37] Brandvoll O. Chemical looping combustion: fuel conversion with inherent [70] Hong J, Kirchen P, Ghoniem AF. Numerical simulation of ion transport
carbon dioxide capture. Department of Energy and Process Engineering, membrane reactors: oxygen permeation and transport and fuel conversion. J
Norwegian University of Science and Technology; Trondheim; 2005 Membr Sci 2012;407–408:71–85.
[38] Naqvi R. Analysis of natural gas-fired power cycles with chemical looping [71] Sarrazin P, Galerie A, Fouletier J, Evans H. Mechanisms of high temperature
combustion for CO2 capture. Department of Energy and Process Engineering, corrosion: a kinetic approach, distributed worldwide by Trans Tech
Norwegian University of Science and Technology; Trondheim; 2006. Publications Limited; 2008.
[39] Kee RJ, Almand BB, Blasi JM, Rosen BL, Hartmann M, Sullivan NP, et al. The [72] Hidayat T, Rhamdhani MA, Jak E, Hayes PC. The kinetics of reduction of dense
design, fabrication, and evaluation of a ceramic counter-flow microchannel synthetic nickel oxide in H2–N2 and H2–H2O atmospheres. Metall Mater Trans
heat exchanger. Appl Therm Eng 2011;31:2004–12. B 2009;40:1–16.
[40] Wang Y, Tonkovich AL, Mazanec T, Daly FP, VanderWiel D, Hu J, et al. Fischer– [73] Garcia-Labiano F, de Diego LF, Adanez J, Abad A, Gayan P. Temperature
Tropsch synthesis using microchannel technology and novel catalyst and variations in the oxygen carrier particles during their reduction and oxidation
microchannel reactor. U.S. Patent No. 7,084,180 [1 August 2006]. in a chemical-looping combustion system. Chem Eng Sci 2005;60:851–62.
Z. Zhao et al. / Fuel 121 (2014) 327–343 343

[74] Wagner C. Beitrag zur Theorie des Anlaufsvorgangs. Z Phys Chem 1933:25–41. [82] Kurkela E, Ståhlberg P. Air gasification of peat, wood and brown coal in a
[75] Mills AF. Heat transfer. 2nd ed. Englewood Cliffs, New Jersey: Prentice Hall; pressurized fluidized-bed reactor. I. Carbon conversion, gas yields and tar
1998. formation. Fuel Process Technol 1992;31:1–21.
[76] Ghoniem AF. Needs, resources and climate change: clean and efficient [83] Milne T, Evans RJ, Abatzoglou N. Biomass gasifier ‘‘tars’’: their nature,
conversion technologies. Prog Energy Combust Sci 2011;37:15–51. formation and conversion (No. TP-570-25357). Boulder, CO: National
[77] Siriwardane R, Poston J, Chaudhari K, Zinn A, Simonyi T, Robinson C. Chemical- Renewable Energy Laboratory (NREL); 1998.
looping combustion of simulated synthesis gas using nickel oxide oxygen [84] Ranzi E. Chemical kinetics of biomass pyrolysis. Energy Fuels
carrier supported on bentonite. Energy Fuels 2007;21:1582–91. 2008;22(6):4292–300.
[78] Jin H, Ishida M. A new type of coal gas fueled chemical-looping combustion. [85] Narváez I, Orío A, Aznar MP, Corella J. Biomass gasification with air in an
Fuel 2004;83:2411–7. atmospheric bubbling fluidized bed. Effect of six operational variables on the
[79] Wu G, Sirignano WA, Williams FA. Simulation of transient convective burning quality of the produced raw gas. Ind Eng Chem Res 1996;35:2110–20.
of an n-octane droplet using a four-step reduced mechanism. Combust Flame [86] Chakravarthy VK, Daw CS, Pihl JA. Thermodynamic analysis of alternative
2011;158:1171–80. approaches to chemical looping combustion. Energy Fuels 2011;25:656–69.
[80] NIST Standard Reference Database Number 69. <webbook.nist.gov/chemistry> [87] McGlashan NR. Chemical-looping combustion – a thermodynamic study. Proc
[accessed 12.2011]. Inst Mech Eng, Part C 2008;222:1005–19.
[81] Sichel EK, Miller RE, Abrahams MS, Buiocchi CJ. Heat capacity and thermal
conductivity of hexagonal pyrolytic boron nitride. Phys. Rev. B 1976;13:
4607–11.

You might also like