16
Kinetic determination of a highly reactive impregnated Fe 2 O 3 /Al 2 O 3 oxygen carrier for use in gas-fueled Chemical Looping Combustion A. Cabello, A. Abad , F. García-Labiano, P. Gayán, L.F. de Diego, J. Adánez Instituto de Carboquímica (ICB-CSIC), Department of Energy and Environment, Miguel Luesma Castán 4, Zaragoza 50018, Spain highlights The reduction and oxidation kinetics of Fe20cAl material was determined by TGA. FeAl 2 O 4 was formed as the only stable Fe- based phase during reduction reactions. The shrinking core model (SCM) was applied to calculate the kinetic parameters. The minimum solids inventory in a CLC unit was 150 kg MW 1 for CH 4 combustion at 1223 K. Reduction reactions were controlled by chemical reaction and diffusion in product layer. graphical abstract Scheme of the reacting mechanism for CH4 reduction with the Fe20cAl oxygen carrier. Al 2 O 3 Al 2 O 3 Al 2 O 3 1. Fresh parcle 2. Kinec control 3. Diffusion control CH 4 CH 4 Fe 2 O 3 Fe 2 O 3 FeAl 2 O 4 CO 2 + H 2 O FeAl 2 O 4 Fe 2 O 3 O CH 4 CO 2 + H 2 O O article info Article history: Received 26 May 2014 Received in revised form 15 July 2014 Accepted 16 July 2014 Available online 25 July 2014 Keywords: CO 2 capture Chemical Looping Combustion Oxygen carrier Iron Reaction kinetic Impregnated abstract The objective of this work was to determine the kinetic parameters for reduction and oxidation reactions of a highly reactive Fe-based oxygen carrier for use in Chemical Looping Combustion (CLC) of gaseous fuels containing CH 4 , CO and/or H 2 , e.g. natural gas, syngas and PSA-off gas. The oxygen carrier was pre- pared by impregnation of iron on alumina. The effect of both the temperature and gas concentration was analyzed in a thermogravimetric analyzer (TGA). The grain model with uniform conversion in the particle and reaction in grains following the shrinking core model (SCM) was used for kinetics determination. It was assumed that the reduction reactions were controlled by two different resistances: the reaction rate was controlled by chemical reaction in a first step, whereas the mechanism that controlled the reactions at higher conversion values was diffusion through the product layer around the grains. Furthermore, it was found that the reduction reaction mech- anism was based on the interaction of Fe 2 O 3 with Al 2 O 3 in presence of the reacting gases to form FeAl 2 O 4 as the only stable Fe-based phase. The reaction order values found for the reducing gases were 0.25, 0.3 and 0.6 for CH 4 ,H 2 and CO, respectively, and the activation energy took values of between 8 kJ mol 1 (for H 2 ) and 66 kJ mol 1 (for CH 4 ). With regard to oxidation kinetics, the reacting model assumed a reaction rate that was only controlled by chemical reaction. Values of 0.9 and 23 kJ mol 1 were found for reaction order and activation energy, respectively. Finally, the solids inventory needed in a CLC system was also estimated by considering kinetic param- eters. The total solids inventory in the CLC unit took a minimum value of 150 kg MW 1 for CH 4 combus- tion, which is a low value when compared to those of other Fe-based materials found in the literature. Ó 2014 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.cej.2014.07.083 1385-8947/Ó 2014 Elsevier B.V. All rights reserved. Abbreviations: CGSM, Changing Grain Size Model; CLC, Chemical Looping Combustion; SCM, shrinking core model; DRM, Diffusion Reaction Model; ICP-AES, inductively coupled plasma atomic emission spectroscopy; MVM, Modified Volumetric Model; TGA, thermogravimetric analyzer; TPR, temperature-programmed reduction; VM, Volumetric Model; XRD, X-ray diffraction. Corresponding author. Tel.: +34 976 733 977; fax: +34 976 733 318. E-mail address: [email protected] (A. Abad). Chemical Engineering Journal 258 (2014) 265–280 Contents lists available at ScienceDirect Chemical Engineering Journal journal homepage: www.elsevier.com/locate/cej

Cabello Et Al

Embed Size (px)

DESCRIPTION

Cabello Et Al

Citation preview

Chemical Engineering Journal 258 (2014) 265–280

Contents lists available at ScienceDirect

Chemical Engineering Journal

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

Kinetic determination of a highly reactive impregnated Fe2O3/Al2O3

oxygen carrier for use in gas-fueled Chemical Looping Combustion

http://dx.doi.org/10.1016/j.cej.2014.07.0831385-8947/� 2014 Elsevier B.V. All rights reserved.

Abbreviations: CGSM, Changing Grain Size Model; CLC, Chemical Looping Combustion; SCM, shrinking core model; DRM, Diffusion Reaction Model; ICP-AES, indcoupled plasma atomic emission spectroscopy; MVM, Modified Volumetric Model; TGA, thermogravimetric analyzer; TPR, temperature-programmed reductVolumetric Model; XRD, X-ray diffraction.⇑ Corresponding author. Tel.: +34 976 733 977; fax: +34 976 733 318.

E-mail address: [email protected] (A. Abad).

A. Cabello, A. Abad ⇑, F. García-Labiano, P. Gayán, L.F. de Diego, J. AdánezInstituto de Carboquímica (ICB-CSIC), Department of Energy and Environment, Miguel Luesma Castán 4, Zaragoza 50018, Spain

h i g h l i g h t s

� The reduction and oxidation kineticsof Fe20cAl material was determinedby TGA.� FeAl2O4 was formed as the only stable Fe-

based phase during reduction reactions.� The shrinking core model (SCM) was

applied to calculate the kineticparameters.� The minimum solids inventory in a CLC

unit was 150 kg MW�1 for CH4

combustion at 1223 K.� Reduction reactions were controlled by

chemical reaction and diffusion inproduct layer.

g r a p h i c a l a b s t r a c t

Scheme of the reacting mechanism for CH4 reduction with the Fe20cAl oxygen carrier.

CH4 CH4CO2 + H2O

CHCO2 + H2O

Al2O3 Al2O3 Al2O3

1. Fresh par�cle 2. Kine�c control 3. Diffusion control

Fe2O3Fe2O3

FeAl2O4 FeAl2O4

Fe2O3

O4

O

a r t i c l e i n f o

Article history:Received 26 May 2014Received in revised form 15 July 2014Accepted 16 July 2014Available online 25 July 2014

Keywords:CO2 captureChemical Looping CombustionOxygen carrierIronReaction kineticImpregnated

a b s t r a c t

The objective of this work was to determine the kinetic parameters for reduction and oxidation reactionsof a highly reactive Fe-based oxygen carrier for use in Chemical Looping Combustion (CLC) of gaseousfuels containing CH4, CO and/or H2, e.g. natural gas, syngas and PSA-off gas. The oxygen carrier was pre-pared by impregnation of iron on alumina. The effect of both the temperature and gas concentration wasanalyzed in a thermogravimetric analyzer (TGA).

The grain model with uniform conversion in the particle and reaction in grains following the shrinkingcore model (SCM) was used for kinetics determination. It was assumed that the reduction reactions werecontrolled by two different resistances: the reaction rate was controlled by chemical reaction in a firststep, whereas the mechanism that controlled the reactions at higher conversion values was diffusionthrough the product layer around the grains. Furthermore, it was found that the reduction reaction mech-anism was based on the interaction of Fe2O3 with Al2O3 in presence of the reacting gases to form FeAl2O4

as the only stable Fe-based phase. The reaction order values found for the reducing gases were 0.25, 0.3and 0.6 for CH4, H2 and CO, respectively, and the activation energy took values of between 8 kJ mol�1 (forH2) and 66 kJ mol�1 (for CH4). With regard to oxidation kinetics, the reacting model assumed a reactionrate that was only controlled by chemical reaction. Values of 0.9 and 23 kJ mol�1 were found for reactionorder and activation energy, respectively.

Finally, the solids inventory needed in a CLC system was also estimated by considering kinetic param-eters. The total solids inventory in the CLC unit took a minimum value of 150 kg MW�1 for CH4 combus-tion, which is a low value when compared to those of other Fe-based materials found in the literature.

� 2014 Elsevier B.V. All rights reserved.

uctivelyion; VM,

Nomenclature

Ci concentration of reacting gas i (mol m�3)�Ci average concentration of gas i in the reactor (mol m�3)Ci,0 inlet concentration of gas i into the reactor (mol m�3)D stoichiometric factor in the fuel combustion reaction

with oxygen (mol O2 per mol fuel)dp particle diameter (lm)Dpl effective product layer diffusivity (m3n0 mol�n0 s�1)Dpl,0 pre-exponential factor for effective product layer diffu-

sivity (m3n0 mol�n0 s�1)Eact activation energy of the kinetic studies presented in

Table 1 (kJ mol�1)Ech activation energy for the chemical reaction (kJ mol�1)Epl activation energy for the diffusion through the product

layer reaction (kJ mol�1)EX activation energy which modifies the decay constant for

the product layer diffusivity (kJ mol�1)Fg shape factor for the grainks chemical kinetic constant (m3n mol�n s�1)ks,0 pre-exponential factor for chemical kinetic constant

(m3n mol�n s�1)kX decay constant for the product layer diffusivitykX,0 pre-exponential factor for decay constant for the prod-

uct layer diffusivitym actual mass of the oxygen carrier (kg)_mc characteristic circulation rate (kg s�1 MW�1)_mOC circulation rate of fully oxidized oxygen carrier (kg s�1

MW�1)mOC total solids inventory in the air reactor and fuel reactor

(kg MW�1)mOC;j solids inventory in the reactor j (kg MW�1)mox mass of the sample of oxygen carrier when it is fully

oxidized (kg)mred mass of the sample of oxygen carrier in reduced form

(kg)MO molecular weight of oxygen (g mol�1)N reaction order for reacting gas in gas–solid chemical

reactionn0

exponential constant for gas concentration in the prod-uct layer diffusion process

Rg constant for ideal gases (Rg = 8.314 J mol�1 K�1)RO oxygen transport capacity of the pure metal oxide (kg

oxygen per kg metal oxide)ROC oxygen transport capacity of the oxygen carrier (kg oxy-

gen per kg solids)R0OC maximum oxygen transport capacity of the oxygen car-

rier when the reaction kinetics is controlled by chemicalreaction

T time (s)T temperature (K)Vg,Xg=0 volume of the gas mixture at Xg = 0 (m3)Vg,Xg=1 volume of the gas mixture at Xg = 1 (m3)xMeO metal oxide contentXg gas conversionXg,in gas conversion at the reactor inlet

Xg,out gas conversion at the reactor outletXS conversion of solidsXS,ox in AR oxidation conversion of solids at the inlet of the air reac-

torX0S,ox in AR normalized oxidation conversion of solids at the inlet of

the air reactor when the reaction kinetics is controlledby chemical reaction

XS,ox in FR oxidation conversion of solids at the inlet of the fuelreactor

X0S,ox in FR normalized oxidation conversion of solids at the fuelreactor inlet when the reaction kinetics is controlledby chemical reaction

XS,ox min minimum oxidation conversion of solids when the reac-tion kinetics is controlled by chemical reaction

XS,red in FR reduction conversion of solids at the inlet of the fuelreactor

X0S,red in FR normalized reduction conversion of solids at the fuelreactor inlet when the reaction kinetics is controlledby chemical reaction

Greek symbolse porosity of particleseg coefficient of expansion of the gas mixtureDH variation of enthalpy for reduction or oxidation reac-

tions (kJ mol�1)DH0

C standard heat of the gas fuel combustion (kJ mol�1)DXg variation of gas conversionDXS variation of solids conversion between fuel reactor and

air reactorDX0S normalized variation of solids conversion between fuel

reactor and air reactor when the reaction kinetics iscontrolled by chemical reaction

DXS,red ch variation of solids conversion between fuel reactor andair reactor when the reaction kinetics is controlled bychemical reaction

sch time for complete conversion when the chemical reac-tion controls the process (s)

so time for complete conversion of particles in the air reac-tor (s)

spl time for complete conversion when the diffusionthrough the product layer controls the process (s)

sr time for complete conversion of particles in the fuelreactor (s)

s0r time for complete conversion of particles in the fuelreactor when the reaction kinetics is only controlledby chemical reaction (s)

Uj characteristic reactivity in the reactor jU0j characteristic reactivity in the reactor j when the reac-

tion kinetics is only controlled by chemical reaction

SubscriptsI gas (CH4, H2, CO)J reactor (AR = air reactor, FR = fuel reactor)Ox oxidationRed reduction

266 A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280

1. Introduction

The Intergovernmental Panel on Climate Change [1] concludedin 2013 that climate change is unequivocal and that its main causeis human activity. In order to avoid an excessive increase in theaverage temperature in the Earth, which could lead to unpredict-able consequences, it is necessary to apply mitigation measures

to reduce worldwide greenhouse gas emissions. Carbon Captureand Storage (CCS) technologies have been proposed as a transitorymeasure until other competitive CO2 emission-free technologiesare demonstrated. CCS is a process consisting of the separation ofCO2 from industrial and energy-related sources, transport to a stor-age location and long-term isolation from the atmosphere [2].According to the IPCC [2], CCS technologies will play a significant

A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280 267

role in reducing CO2 emissions and stabilizing its concentration inthe atmosphere with reasonable cost-effectiveness, mitigating itsnegative effects on the environment. The main disadvantages ofthese technologies are their higher initial costs and the energy pen-alty on the overall combustion process. In this context, over thelast years there has been considerable growth in interest in devel-oping new, cost-effective combustion processes that generatehighly concentrated CO2 streams. One of these new processes isChemical Looping Combustion (CLC) which has emerged as anenvironmentally-acceptable alternative for energy productionfrom fossil fuels given that CO2 separation is inherent to the actualcombustion process.

The concept on which the CLC process is based was first pro-posed in 1954 by Lewis and Gilligand [3] to produce pure CO2 fromfossil fuels. Three decades later, in 1983, Ritcher and Knoche [4]presented CLC technology as a suitable process with which toincrease the thermal efficiency of a power plant, which was sup-ported by works performed by Ishida et al. [5–7]. At the beginningof the current century Lyngfelt et al. [8] proposed CLC for the cap-ture CO2 at a low cost. The CLC design was based on two intercon-nected fluidized beds with a solid, known as an oxygen carrier,circulating between them. This oxygen carrier is often composedof a metal oxide, which is the source of oxygen in this process,and an inert material which acts as support in order to increasemechanical strength. The main role of this oxygen carrier is to pre-vent the direct contact between fuel and air during the combustionprocess, for which this material is circulated between two reactors:the fuel reactor, where the reduction of the oxygen carrier andcombustion of the fuel take place, and the air reactor, where theoxygen carrier is oxidized.

One key aspect for optimum performance of a CLC system is theoxygen carrier, which has to be able to fully convert the fuel to CO2

and H2O and to present high reactivity with the fuel and airthroughout a large number of redox cycles. Furthermore, it mustfulfill other characteristics, such as high attrition resistance andnot presenting any problems of agglomeration or carbon deposi-tion. Finally, environmental and economic aspects must be alsoconsidered for the final selection of an optimum oxygen carrier [9].

Iron-based oxygen carriers are gaining considerable importancefor the CLC process, since they present better environmental com-patibility and a lower cost in comparison with other metal oxides.These oxygen carriers have shown adequate reactivity underatmospheric [10–12] and pressurized [13] conditions and highreactivity with CO and H2, and do not present thermodynamic lim-itations to fully convert CH4, CO and H2 to CO2 and H2O if thereduction of the iron oxide, in the form of haematite (Fe2O3), is lim-ited to the form of magnetite (Fe3O4). Nevertheless, if the Fe-basedoxygen carriers include Al2O3 as support material, FeAl2O4 can beformed as a reduced compound and also achieve complete com-bustion of gas to CO2 and H2O [14–17]. Thus, the oxygen transportcapacity of the oxygen carrier is increased three times in compar-ison with the Fe2O3–Fe3O4 redox couple. Furthermore, Fe-basedoxygen carriers present a low tendency to carbon deposition [18]and no risk of sulfide or sulfate formation at any concentrationor operating temperature when H2S-containing gases are used asfuels [17,19,20].

Fe-based oxygen carriers have traditionally presented lowerreactivity with fuel gases compared to Ni-, Cu-, and Mn-based oxy-gen carriers, mainly for reduction with CH4. Consequently, incom-plete combustion of CH4 occurred in different continuous CLCplants when Fe-based materials were used [21–28]. Among thesematerials there were synthetic oxygen carriers prepared by differ-ent methods [21–23], iron-containing minerals [24–27] and resi-dues [27,28]. Our research group at Instituto de Carboquímica(ICB-CSIC) has recently developed a highly reactive Fe-basedoxygen carrier prepared by the impregnation method [16]. This

material showed complete combustion of CH4 during continuousoperation in a 0.5 kWth CLC unit [16]. Additionally, it presentedadequate resistance to agglomeration and attrition, even whenH2S was present in the gas fuel at very high concentrations[17,19,20].

Process modeling is a powerful tool with which to design andoptimize a CLC unit. A proper model of the fuel and air reactorsin a CLC system must take into consideration the oxygen transportcapacity and the reaction kinetics of the oxygen carrier particles.Kinetic determination of redox reactions becomes crucial for thesimulation, design and optimization of CLC units because the solidsinventory needed in the fuel and air reactors is directly related tooxygen carrier reactivity [29]. Moreover, oxygen transport capacitygreatly affects the solids recirculation rate needed between the fueland the air reactors.

Several works deal with the kinetic determination of reductionand oxidation reactions between oxygen carriers and commongases used for CLC, i.e., CH4, CO, H2 and O2. Table 1 shows a sum-mary of the research works found in the literature regardingkinetic data determined for Fe-based oxygen carrier particles. Thistable includes information related to the composition of the oxy-gen carriers, method of preparation, properties of the evaluatedmaterials, experimental conditions studied, and the type of reactorand model used to determine the kinetic data. The existence of dif-ferent conditions inside the fluidized beds makes it essential todetermine the reaction rates for a wide range of temperature andfuel concentration values. Furthermore, it is also important to cal-culate kinetic data for all relevant gases, i.e. CH4, CO, H2 and O2, andFe-based redox systems applicable for CLC process, such as Fe2O3–Fe3O4, Fe2O3–FeAl2O4 and Fe2TiO5–FeTiO3. From the analysis of thedata collected in Table 1, it can be concluded that most of theworks produced to date do not provide sufficient information withwhich to model a CLC system using an Fe-based material as anoxygen carrier. Theoretical kinetic studies, relevant from a basic-science point of view, have been also conducted to investigatethe mechanisms of interaction between reacting gases and Fe-based oxygen carriers [39–48], suggesting that the Fe2O3-supportinteraction influences material reactivity. However, these worksdo not provide enough practical information for CLC system designand simulation.

The aforementioned Fe-based material [16] can be considered apromising oxygen carrier for use during the scale-up of the CLCprocess. Therefore, determination of kinetic data would be essen-tial in order to design, optimize and scale-up a CLC system basedon this oxygen carrier. CH4, CO and/or H2 are considered majorcomponents in natural gas, syngas and PSA-off gas. Thus, the objec-tive of this work was to determine the kinetic parameters for thereduction of the highly reactive Fe-based oxygen carrier withCH4, CO and/or H2, as well as the kinetics of oxidation with O2.The effect of both temperature and gas concentration on reactionrate was evaluated. Moreover, the obtained parameters werefurther used to estimate the recirculation rate and the solids inven-tory needed in a CLC system.

2. Material and methods

2.1. Oxygen carrier

A kinetic study of redox reactions for CLC was conducted for apromising Fe-based oxygen carrier prepared at ICB-CSIC. This oxy-gen carrier was prepared by the incipient impregnation methodand was composed of Fe2O3, as active material, supported on c-Al2O3. A detailed description of the method of preparation can befound elsewhere [16]. The total Fe2O3 content in the particles usedin this work was 20 wt.%. The main physicochemical properties of

Table 1Summary of kinetic data for Fe-based oxygen carriers.

Oxygen carrier Method of preparation Experimentalconditions

Kinetic model Redox system Reference

Fe2O3

dp = 125 lmROC = 10.0%

Solid state calcination TGA; T = 773–1173 KCH4

Diffusion controln = n.a.; Eact = 271 kJ/mol

Fe2O3/FeO [30]

Fe2O3

dp = 300–425 lmROC = 3.3–10.0%

Mechanical mixing Fluid bed; T = 723–1123 K1.5–10 vol.% CO

DRMn = 1.0; Eact = 75 kJ/moln = 1.0; Eact = 94 kJ/mol

Fe2O3/Fe3O4

Fe3O4/FeO[31]

Multi-crystal Fedp = n.a.ROC = 30.0%

n.a. TGA; T = 973–1173 K0.07–5.25 vol.% O2

Nucleation and growth modeln = n.a.; Eact = n.a.

Fe2O3/Fe [32]

94 wt.% Fe2O3

dp = 60–100 lmROC = 9.4%

Crush and sieve TGA; T = 973–1098 K15–35 vol.% CH4

Two parallel reactions (R1 and R2)R1: n = n.a.; Eact = 34 kJ/molR2: nucleation and grow model;n = n.a.; Eact = 39 kJ/mol

Fe3O4/FeOFe2O3/Fe3O4

[33]

45 wt.% Fe2O3 on Al2O3

dp = 90–250 lmROC = 1.3%

Impregnation TGA; T = 773–1273 KTotal pressure: 1–30 atm5–70 vol.% H2

5–70 vol.% CO5–21 vol.% O2

CGSMn = 0.8; Eact = 24 kJ/moln = 1.0; Eact = 20 kJ/moln = 1.0; Eact = 14 kJ/mol

Fe2O3/Fe3O4 [13]

58 wt.% Fe2O3 on Al2O3

dp = 90–106 lmROC = 4.0%

Mechanical mixing TGA; T = 1073–1123 K20–70 vol.% CH4

20–70 vol.% H2

20–70 vol.% CO

SCM with Fg = 3n = 0.2; Eact = 45 kJ/moln = 0.85; Eact = 22 kJ/moln = 1.0; Eact = 19 kJ/mol

Fe2O3/Fe3O4–FeAl2O4

[34]

60 wt.% Fe2O3 on Al2O3

dp = 90–250 lmROC = 4.1%

Freeze granulation TGA; T = 873–1223 K5–70 vol.% CH4

5–70 vol.% H2

5–70 vol.% CO5–21 vol.% O2

SCM with Fg = 3n = 1.3; Eact = 49 kJ/moln = 0.5; Eact = 24 kJ/moln = 1.0; Eact = 20 kJ/moln = 1.0; Eact = 14 kJ/mol

Fe2O3/Fe3O4–FeAl2O4

[14,29]

60 wt.% Fe2O3 on bentonitedp = 106–150 lmROC = 2.0%

Mechanical mixing TGA; T = 973–1273 K10 vol.% CH4

10 vol.% O2

Red: MVM; Ox: SCMn = n.a.; Eact = 29 kJ/moln = n.a.; Eact = 6 kJ/mol

Fe2O3/Fe3O4 [21]

75–95 wt.% Fe2O3 on MgOdp = 100–300 lmROC = 7.5–9.5%

Hot incipient wetnessimpregnation

TGA; T = 973–1098 K5–20 vol.% CH4

Two-competing reactions(R1 and R2)R1: n = n.a; Eact = 50 kJ/molR2: n = n.a; Eact = 65 kJ/mol

Fe2O3/Fe3O4

Fe3O4/FeO[35]

80 wt.% Fe2O3 on TiO2

dp < 250 lmROC = 2.7–24.0%

Solid-state mixing TGA; T = 873–1223 K3 vol.% H2 in ArAir

Red:VM; Ox: SCMn = n.a.; Eact = 34 kJ/moln = n.a.; Eact = 0 kJ/mol

Fe2O3/Fe3O4–FeO–Fe

[36]

Pre-oxidized ilmenite(Fe2TiO5)dp = 150–300 lmROC = 4.0%

Thermal treatment TGA; T = 1073–1123 K5–50 vol.% CH4

5–50 vol.% H2

5–50 vol.% CO5–21 vol.% O2

SCM with Fg = 3n = 1.0; Eact = 165 kJ/moln = 1.0; Eact = 109 kJ/moln = 1.0; Eact = 113 kJ/moln = 1.0; Eact = 12 kJ/mol

Fe2O3�TiO2/FeO�TiO2

[37]

Activated ilmenite (Fe2TiO5)dp = 150–300 lmROC = 3.3%

Thermal treatment +activation in a fluidizedbed

TGA; T = 1073–1123 K5–50 vol.% CH4

5–50 vol.% H2

5–50 vol.% CO5–21 vol.% O2

SCM with Fg = 3n = 1.0; Eact = 136 kJ/moln = 1.0; Eact = 65 kJ/moln = 0.8; Eact = 80 kJ/moln = 1.0; Eact = 25 kJ/mol

Fe2O3�TiO2/FeO�TiO2

[37]

98 wt.% Fe2O3

dp = 50–200 lmROC = 3.2%

Crush and sieve TGA; T = 1073–1223 K33 vol.% CH4

Avrami–Erofe’ev phase changemodeln = n.a.; Eact = 215 kJ/mol

Fe2O3/Fe3O4 [38]

268 A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280

the oxygen carrier particles are shown in Table 2. This oxygen car-rier is hereafter referred to as Fe20cAl.

2.2. Characterization techniques

Different techniques were used to physically and chemicallycharacterize fresh particles of the Fe20cAl material. The totalFe2O3 content was determined by inductively coupled plasmaatomic emission spectroscopy (ICP-AES) with a Jobin Ybon 2000spectrometer. The mean particle size was measured by means ofsize particle distribution, via laser diffraction technique, accordingto the ISO 13320 Standard with LS 13 320 Beckman Coulter equip-ment. The skeletal density of the particles was determined with a

Micromeritics Model AccuPyc II 1340 helium pycnometer. AShimpo FGN-5X apparatus was used to measure the crushingstrength of particles as the force needed to fracture them. Crushingstrength was obtained as the average value of at least 20 measure-ments for particles with a particle size range of dp = +200–400 lm.Porosity was measured by Hg intrusion in a Quantachrome Pore-Master 33, whereas the specific surface was determined by theBrunauer–Emmett–Teller (BET) method by adsorption/desorptionof nitrogen at 77 K in a Micromeritics ASAP-2020 (MicromeriticsInstruments Inc.). Crystalline chemical species were identified bypowder X-ray diffraction (XRD) in a Bruker AXS D8 Advance sys-tem, with Bragg–Brentano geometry configuration, Cu Ka radiationand equipped with secondary graphite monochromator. Finally,

Table 2Main physicochemical properties of the Fe20cAl material.

Fresh material

Fe2O3 (wt.%) 201

Oxygen transport capacity, ROC 0.02Particle size (lm) 200–400Skeletal density (kg m�3) 3950Crushing strength (N) 1.5Porosity (%) 50.5Specific surface area, BET (m2 g�1) 39.1XRD Fe2O3, a-Al2O3

1 Determined by ICP-AES.

A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280 269

the reducibility of the Fe-based oxygen carrier particles was deter-mined by temperature-programmed reduction (TPR) experimentsin an AUTOCHEM II flow apparatus from Micromeritics. In theTPR technique, the temperature was increased from room temper-ature to 1273 K at a constant rate of 7 K min�1 with a flow of20 ml min�1 of a 10 vol.% H2/90 vol.% Ar mixture.

2.3. Experimental setup and procedure

The kinetics of reduction and oxidation reactions with the Fe-based oxygen carrier was determined from experiments at atmo-spheric pressure in a TGA, CI Electronics type, described elsewhere[37]. Three different gases were used as reducing agents, i.e., CH4,CO and H2, whereas O2 was used for the oxidation reactions. In allcases, the redox agents were diluted in N2.

The samples of approximately 50 mg of oxygen carrier wereloaded in a platinum basket, and a gas flow of 25 LN/h was fed intothe reactor of the TGA system. External and interparticle diffusionhave low relevance under these conditions [14]. Furthermore,García-Labiano et al. [49] found that oxygen carrier particles canbe considered isotherm under usual operating conditions, and thatmass transfer inside them is negligible.

The oxygen carrier particles were heated to the operating tem-perature in an air atmosphere. When the sample weight was stabi-lized, the oxygen carrier was exposed to a reduction and anoxidation cycle. In order to prevent the mixing of fuel and air, a2-min flow of N2 was fed between the reducing and oxidizing per-iod. This material showed excellent stability in redox reactivityover hundreds of cycles [17]. Data obtained during the first redoxcycle were selected for kinetic determination.

Different fuel gas (CH4, CO and H2) concentrations ranging from5 to 60 vol.% were used to determine the kinetics of the reductionreaction. In order to prevent carbon deposition, 20 vol.% of H2O orCO2 were added to the reacting mixture when CH4 or CO were usedas fuel gases, respectively. To analyze oxidation reactivity, it wasinitially necessary to reduce the sample to FeAl2O4, simulatingthe behavior expected in a CLC process when iron oxide (Fe2O3)supported on Al2O3 is considered as the oxygen carrier. In this case,as mentioned in the introduction section, it was possible to reachcomplete fuel combustion to CO2 and H2O, obtaining FeAl2O4 asthe reduced form of the oxygen carrier. For this reason, the previ-ous reducing periods were conducted at 1223 K in a 15 vol.%CO + 20 vol.% CO2 (N2 to balance) atmosphere. Under these condi-tions and with the addition of CO2, the reduction of Fe2O3 wasstopped at FeAl2O4, and no further reduction to metallic iron wasproduced. This fact was also confirmed by XRD analysis of reducedsamples. The oxygen carrier particles were subsequently ready tobe subjected to the oxidizing period. The O2 concentration for oxi-dation was varied from 5 to 21 vol.%.

To study the effect of the temperature on reduction and oxida-tion reaction rates, this parameter was varied between 973 and1323 K. When the reduction reaction was studied, the fuel gas

concentration was always 15 vol.% and the samples were oxidizedat 1223 K in an air atmosphere. On other hand, in order to evaluatethe influence of temperature on oxidation kinetics, a 10 vol.% O2

concentration was always used and, as mentioned above, thereducing period was conducted at 1223 K in a 15 vol.%CO + 20 vol.% CO2 (N2 to balance) atmosphere.

2.4. Data evaluation

The following reactions took place between Fe2O3 and thereducing gases:

4Fe2O3 þ 8c-Al2O3 þ CH4 () 8FeAl2O4 þ 2H2Oþ CO2

DH1223Kred ¼ �348:7 kJ=mol ð1Þ

Fe2O3 þ 2c-Al2O3 þH2 () 2FeAl2O4 þH2O

DH1223Kred ¼ �135:5 kJ=mol ð2Þ

Fe2O3 þ 2c-Al2O3 þ CO() 2FeAl2O4 þ CO2

DH1223Kred ¼ �168:1 kJ=mol ð3Þ

As can be observed, all the reduction reactions were exothermicat 1223 K, a usual operating temperature for the CLC process.

For the oxidation of FeAl2O4, the following reaction wasexpected:

4FeAl2O4 þ O2 () 2Fe2O3 þ 4c-Al2O3 DH1223Kox ¼ �227:1 kJ=mol

ð4Þ

The conversion of solids, XS, was calculated with Eqs. (5) and (6)for the reduction or oxidation period, respectively.

XS;red ¼mox �m

mox �mred¼ mox �m

ROC �moxð5Þ

XS;ox ¼m�mred

mox �mred¼ 1� mox �m

ROC �moxð6Þ

where m is the actual mass of the sample, mred the mass ofreduced sample, mox the mass of oxidized sample. The differencebetween mox and mred is the maximum amount of oxygen thatcan be transferred from the oxygen carrier to the fuel, which isdefined by the oxygen transport capacity, ROC. In the case of ironmaterials, this amount of oxygen depends on the final oxidationstate during the reduction reaction. In this kinetic study, theamount of oxygen that could be provided by the oxygen carriermaterial to allow complete combustion of fuel to CO2 and H2Owas considered to correspond to the reduction from hematite(Fe2O3) to iron aluminate (FeAl2O4) [9]. Therefore, the conversionof solids and, consequently, the reduction and oxidation kineticswere calculated by taking this redox couple as a reference.

3. Results

3.1. Analysis of Fe2O3 reduction to FeAl2O4

Prior to conducting the kinetic study of this oxygen carrier, eval-uation was made of what would be the most appropriate methodfor kinetic determination. There are mainly two methods for kinet-ics determination, known as the isothermal method and non-iso-thermal method. In this sense, different experiments forreduction reaction were carried out in a TGA system either at a fixedtemperature or by TPR. From a TPR profile, see Fig. 1, three hydro-gen consumption peaks were well identified at 683, 985 and1132 K, respectively. These three peaks corresponded to the follow-ing reducing transitions: Fe2O3 ? Fe3O4 ? FeO ? Fe [17]. None ofthese peaks could be attributed to the reduction of

270 A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280

FeAl2O4 since the peak corresponding to the reduction from FeAl2O4

to Fe0 would have been detected at a higher temperature [17].In order to analyze the reduction pathway under relevant con-

ditions in CLC, isothermal experiments were conducted at 1223 Kwith a reducing gas mixture of 5 vol.% H2, 48 vol.% H2O (N2 to bal-ance). The reduction period was stopped at different times duringthe operation in order to obtain several degrees of solid conversion.The corresponding samples were further subjected to XRD analyseswith the aim of observing the crystalline iron phases at differentdegrees of oxygen carrier reduction. In all cases, the XRD techniquerevealed that the only Fe-based reduced phase found in the oxygencarrier was iron aluminate, FeAl2O4 together with unreacted Fe2O3,regardless of the degree of the oxygen carrier reduction. This resultwas also corroborated by thermodynamic analyses carried outwith HSC Chemistry 6.1 software [50]. Moreover, tests performedwith this material in a continuous CLC unit for CH4 combustionrevealed the presence of FeAl2O4 in the oxygen carrier particlesextracted from the fuel reactor as the only reduced compound[16,17].

With the above experimental data, it can be concluded that dif-ferent results regarding the reduction mechanism were founddepending on the method selected for kinetic determination. How-ever, the isothermal method, based on the experiments carried outin the TGA at a fixed temperature, provided a better fit with theconditions found in the CLC process. In this respect, it was con-cluded that the isothermal method was suitable to determine reac-tion kinetics of this material for use under typical conditions inCLC.

It was also concluded that the reduction reaction mechanismwas based on the interaction of Fe2O3 with Al2O3 in presence ofthe reacting gases to form FeAl2O4 as the only stable Fe-basedphase. Thus, as previously mentioned, the conversion of solidsand, consequently, the reduction and oxidation kinetics werecalculated taking the Fe2O3(Al2O3)–FeAl2O4 redox couple as areference.

3.2. Isothermal analysis of reduction reaction

The reduction kinetics of the Fe20cAl oxygen carrier was stud-ied in a TGA by using three different fuels: CH4, H2 and CO. Thekinetics were determined in a wide range of temperatures (from973 to 1323 K) and gas concentrations (from 5 vol.% to 60 vol.%)typically found in a CLC unit.

Temperature (K)

400 500 600 700 800 900 1000 1100 1200 1300

H2

con

cen

trat

ion

(vo

l. %

)

8.5

9.0

9.5

10.0

Fig. 1. TPR profile of fresh particles of the Fe20cAl oxygen carrier.

3.2.1. Effect of fuel typeFig. 2 shows the conversion vs time curves for the three fuel

gases used to determine reduction kinetics. The reduction testswith H2 and CH4 were performed by adding 20 vol.% H2O to thereacting mixtures, whereas 20 vol.% of CO2 was added for TGA testswith CO. As can be observed, the reduction rate was very high inthe three cases up to a solid conversion value of 0.6. Beyond thispoint, the oxygen carrier presented a slightly different behaviordepending on the reducing gas being considered. In the cases ofH2 and CH4, the reaction rate decreased rapidly, continuing at avery slow rate from this point. Consequently, complete reductionfrom Fe2O3 to FeAl2O4 was not achieved using these reactingagents. However, when CO was considered as the reducing gas,the reaction rate was also considerably reduced at conversion val-ues higher than 0.6, but it was still high enough to completelyreduce the oxygen carrier to FeAl2O4 after 600 s. At that point,the reduction reaction stopped and no further reduction to metalliciron took place, even though the experiment in the TGA was pro-longed until 3600 s. A remarkable result obtained from the reduc-tion tests with CO is that the active iron oxide content calculatedfrom the TGA tests coincides accurately with the theoretical con-tent derived from the preparation method for the oxygen carrier,i.e. the incipient impregnation method, as well as with the corre-sponding content determined from ICP-AES analysis, see Table 2.The haematite (Fe2O3) content of the oxygen carrier particles was20 wt.% in all three cases. These results mean that reduction testsin a TGA using a mixture of CO and CO2 as reacting gas can beconsidered an additional, valid method to determine the masspercentage of active iron oxide in an Fe-based oxygen carrier sup-ported on alumina. Reduction with H2 gave a value of 15 wt.%Fe2O3 [16], which corresponded to the fraction of Fe2O3 that washighly reactive.

3.2.2. Effect of gas products on the reduction reactionDifferent gas products can be formed during the combustion

process of a gaseous fuel, of which CO2 and H2O are the most rel-evant. The presence of different amounts of these products onthe reaction atmosphere could affect the reaction rate. In orderto evaluate the effect of H2O on the reduction reaction rate of theFe20cAl oxygen carrier, several tests were carried out with H2 act-ing as reducing gas and by varying the H2O/H2 ratio from 0 to 10approximately. Fig. 3a illustrates the reduction conversionobtained for this material with H2 when the H2O content was

Time (s)

0 100 200 300 400 500 600

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

1.0

CH4

CO

H2

Time (s)

0 5 10 15 20 25 30

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

Fig. 2. Conversion vs time curves during reduction period with CH4, CO and H2.Testing conditions: T = 1223 K; reducing gas mixtures: 15 vol.% H2 + 20 vol.% H2O;15 vol.% CO + 20 vol.% CO2; and 15 vol.% CH4 + 20 vol.% H2O (N2 to balance).

A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280 271

varied from 0 to 48 vol.%. When H2O was not added to the reducinggas mixture, the oxygen carrier was reduced beyond FeAl2O4,although it could not be completely reduced to metallic iron. WhenH2O was added to the reducing gas, the reaction rate stopped after50 s of reduction at a solids conversion of 0.75 approximately. Afurther increase in the amount of H2O slightly changed the oxygencarrier reactivity and the reaction rate with H2. This means that thereduction of FeAl2O4 to Fe0 is prevented if H2O is present in thereducing gas mixture.

On the other hand, Fig. 3b shows the conversion vs time curvesobtained with 5 vol.% CO when the CO2 concentration wasincreased from 0 to 50 vol.%. In all cases, the reaction rate was veryfast up to a solids conversion value of 0.75 approximately. Thisdegree of conversion was obtained after 30 s of reduction reaction.For higher times, the reaction rate was considerably decreased, butthe oxygen carrier particles were able to be reduced completely toFeAl2O4. When CO2 was not added to the reducing gas mixture, theoxygen carrier was reduced beyond FeAl2O4. When CO2 was addedtogether with the reducing gas, the behavior of the oxygen carrierwas practically the same, regardless of the CO2 concentration used.Thus, the reduction reactivity with CO was hardly affected whenthe amount of CO2 was varied from 10 to 50 vol.% in the reactiongas mixture.

From the results obtained in this section it can be concludedthat the presence of different concentrations of H2O and CO2, the

Time (s)

0 50 100 150 200 250 300

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

1.0

0 vol.% H2O

10 vol.% H2O

20 vol.% H2O

30 vol.% H2O

40 vol.% H2O

48 vol.% H2O

(a)

Time (s)

0 20 40 60 80 100 120

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

1.0

0 vol.% CO2

10 vol.% CO2

20 vol.% CO2

30 vol.% CO2

40 vol.% CO2

50 vol.% CO2

(b)

Fig. 3. Effect of H2O and CO2 presence on the reduction conversion of the Fe20cAloxygen carrier with H2 (a) and CO (b). Testing conditions: T = 1223 K; reducinggases: 5 vol.% H2 and 5 vol.% CO.

main products of the combustion reaction, in the reacting gas mix-tures affected neither the reaction rate nor the reactivity of theoxygen carrier. Only the avoidance of these gases in the reactiongas mixture increased the reaction rate during the reaction at highsolids conversion values, characterized by slow reactivity. How-ever, H2O and CO2 are considered to be in the gas reacting in thefuel reactor because they would be quickly generated in the fuelreactor, or can even be present in fluidizing gases, e.g. in the fuelgas itself or in gases introduced into loop seals. Therefore, kineticdetermination was obtained considering the presence of H2O orCO2 in the reacting gas mixture.

3.2.3. Effect of gas concentrationThe concentration of fuel gases inside a fuel reactor varies with

height along the fluidized bed. The oxygen carrier is usually in con-tact with high fuel concentration values at the bottom, whereas theconcentration of fuel gases decreases dramatically at the top of thebed in such a way that the gaseous mixture is mainly composed ofCO2 and H2O.

The effect of fuel gas concentration on the reduction of theFe20cAl oxygen carrier was determined by carrying out experi-ments in a TGA with different concentrations of CH4, CO and H2

at a temperature of 1223 K. In the cases of CH4 and H2, the fuelgas concentration was varied from 5 to 30 vol.%, while the concen-tration of CO was increased up to 60 vol.%. The conversion vstime curves corresponding to the three fuel gases are shown inFig. 4(a–c). An increase in the reaction rate was noticed duringthe first part of the reduction period as the fuel gas concentrationwas increased. At the beginning of the period, the reaction rate wasvery fast for all the fuel gases (for 10 s for CH4 and H2 and for 20 sfor CO, respectively), however it decreased immediately, contin-uing at a very low rate for the rest of the period. In this second per-iod, characterized by a low reaction rate, the fuel gas concentrationbarely affected the reduction rate of the oxygen carrier with thedifferent fuel gases. This behavior indicates that the correspondingreduction reactions are controlled by two different resistances. Thechange in controlling mechanism happens at conversion values of0.45 for CH4 and H2 and at 0.7 for CO.

3.2.4. Effect of temperatureThe effect of temperature on the reduction reactivity of the

Fe20cAl oxygen carrier was also evaluated. In this case, the con-centration of the reducing gases for these tests was set at a fixedvalue and the temperature was varied from 973 to 1323 K. Thisrange of values includes the usual operating temperatures in aCLC system. Fig. 5(a–d) shows solids conversion as a function oftime for the following reducing gas mixtures: (a) 15 vol.%CH4 + 20 vol.% H2O; (b) 15 vol.% CO + 20 vol.% CO2; (c) 15 vol.%H2 + 20 vol.% H2O; and (d) 5 vol.% H2 + 48 vol.% H2O. The gas mix-tures were completed with N2 in all four cases. The reaction rateand final reduction conversion of the oxygen carrier were affectedby the temperature, since an increase in this parameter producedan increase in both terms. In the case of the final conversionreached for the oxygen carrier this means that the reductiondegree to iron aluminate (FeAl2O4) was clearly influenced by thetemperature.

Solids conversion was quite low at T < 1073 K, mainly for CH4

and H2, but the reaction rate was relatively fast during the first sec-onds of the reduction period when XS,red < 0.2. As in the case of thestudy of the effect of gas concentration (see Section 3.2.3), thisbehavior indicates that the reducing reaction is controlled by twodifferent resistances depending on solids conversion. At low XS,red

values, the reaction rate was fast. However, as the oxygen carrierparticles were reduced to a greater extent, the control step changesand the reaction rate quickly decreased. The same behavior wasobserved at higher temperatures for all the reducing gases, with

Time (s)

0 10 20 30 40 50 60

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

1.0

5 vol.% CH4

10 vol.% CH4

15 vol.% CH4

20 vol.% CH4

30 vol.% CH4

(a)

Time (s)

0 10 20 30 40 50 60

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

1.0

5 vol.% H2

10 vol.% H2

15 vol.% H2

20 vol.% H2

30 vol.% H2

(b)

Time (s)

0 20 40 60 80 100 120

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

1.0

5 vol.% CO

15 vol.% CO

30 vol.% CO

60 vol.% CO

(c)

Fig. 4. Effect of fuel gas concentration on the reduction reaction for CH4 (a), H2 (b)and CO (c). T = 1223 K. The continuous lines are results predicted by the modelusing kinetic parameters obtained in this work.

272 A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280

the difference that the reduction conversion at which the reactionrate changed from fast to slow increased with temperature. Thisfact was the main reason for the increase in conversion achievedat long times with temperature. The effect of temperature on thereaction rate during the second step, which was characterized bya low reaction rate, was low. This fact can be seen because the con-version vs time curves obtained at different temperatures werepractically parallel.

3.3. Isothermal analysis of oxidation reaction

Oxidation of the oxygen carrier takes place in the air reactor.Fully or partially reduced particles coming from the fuel reactorare exposed to air in the air reactor with the aim of regeneratingthe material for a new reduction step. Different O2 concentrationsexist in the air reactor. The O2 concentration varies along the bedheight from 21 vol.% to the O2 concentration in the exhaust airstream. An O2 concentration of 4 vol.% at the top of the bed, a typ-ical value in a CLC plant [51], is equivalent to 20% excess air in theair reactor. Therefore, the concentration in the studies using theTGA was varied between 5 and 21 vol.% O2 (N2 to balance), andthe temperature from 1073 to 1273 K. The experiments were car-ried out following the procedure described in the experimentalsection.

3.3.1. Effect of O2 concentrationFig. 6 illustrates the conversion of the oxygen carrier during oxi-

dation at 1223 K with 5, 10, 15 and 21 vol.% O2. The oxidation reac-tion rate increased as the O2 concentration was increased. Theseconversion vs time curves evidence the high speed at which theoxidation reaction took place. The oxygen carrier samples werefully oxidized in all cases and the time for complete conversionranged between 15 s and 50 s.

3.3.2. Effect of temperature on the oxidation reactionThe effect of the temperature on the oxidation reaction was also

investigated in this work. The influence of this parameter wasstudied between 1073 and 1273 K. Fig. 7 shows solids conversionas a function of time for an O2 concentration of 10 vol.%. In this casethe variation in temperature scarcely affects the reaction rate. As inthe previous case (see Section 3.2.1), the oxidation reaction wasvery fast and the samples were always fully oxidized at the endof the oxidation period.

3.4. Kinetic model

The kinetic parameters of the reduction and oxidation reactionscan be determined using a particle reaction model that fits ade-quately with the experimental results. For materials prepared byimpregnation, the grain model with uniform conversion in the par-ticle and reaction in grains following shrinking core model [52].From previous figures (Figs. 3–7), some information was acquiredregarding the mechanisms controlling the reduction and oxidationreactions with the Fe20cAl oxygen carrier. The reaction rate wasobserved to be very fast with all gases at the beginning of thereduction period. However, it decreased immediately and contin-ued at a very low rate for the rest of the period. This behavior indi-cates that the corresponding reduction reactions are controlled bytwo different resistances. Furthermore, it should be taken intoaccount that magnetite (Fe3O4) never appears as the Fe-basedreduced phase and FeAl2O4 was the only reduced compound inthe presence of H2O or CO2 when the oxygen carrier was reducedby H2 or CO2, respectively.

Fig. 8, is a schematic illustration of the proposed reacting mech-anism for CH4 reduction with the Fe20cAl oxygen carrier. Thereacting model for reduction kinetics assumes a first step withthe reaction rate controlled by chemical reaction in the grainsurface [53,54]. During this first step, alumina and oxygen mustdiffuse towards the reaction interphase. As a result, a product layerof FeAl2O4 is formed. In the second step, the mechanism that con-trols the reaction at higher conversion values is the diffusionthrough the product layer of FeAl2O4 around the grains. Moreover,the reduction rate during this step is independent of the fuel gasconcentration. This fact suggests that diffusion of the reacting gasis blocked, but oxygen must diffuse outwards [53]. Thus, chemical

Time (s)

0 10 20 30 40 50 60

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

1.0

(a)15 vol.% CH41073 K

1123 K

1173 K

1223 K

1273 K

Time (s)

0 20 40 60 80 100 120

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

1.0

(b)15 vol.% CO

1073 K1123 K 1073 K1223 K

Time (s)

0 10 20 30 40 50 60

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

1.0

(c)15 vol.% H2

1073 K

1123 K

1173 K

1223 K

1273 K

Time (s)

0 10 20 30 40 50 60

Co

nve

rsio

n, X

s,re

d

0.0

0.2

0.4

0.6

0.8

1.0

(d)5 vol.% H2973 K

1073 K

1123 K

1173 K

1223 K

1273 K

1323 K

Fig. 5. Effect of the temperature on the reduction reaction of the Fe20cAl oxygen carrier. Reducing gas mixtures: (a) 15 vol.% CH4 + 20 vol.% H2O; (b) 15 vol.% CO + 20 vol.%CO2; (c) 15 vol.% H2 + 20 vol.% H2O; and (d) 5 vol.% H2 + 48 vol.% H2O. The continuous lines are results predicted by the model using kinetic parameters obtained in this work.

Time (s)

0 10 20 30 40 50 60

Co

nve

rsio

n, X

s,o

x

0.0

0.2

0.4

0.6

0.8

1.0

5 vol.% O2

10 vol.% O2

15 vol.% O2

21 vol.% O2

Fig. 6. Effect of O2 concentration on the oxidation reaction of the Fe20cAl oxygencarrier. Operating conditions: T = 1223 K. The continuous lines are results predictedby the model using kinetic parameters obtained in this work.

Time (s)

0 10 20 30 40 50 60

Co

nve

rsio

n, X

s,o

x

0.0

0.2

0.4

0.6

0.8

1.0

1073 K

1123 K

1173 K

1223 K

1273 K

Fig. 7. Effect of temperature on the oxidation reaction of the Fe20cAl oxygen carrierwith O2 (10 vol.%). The continuous lines are results predicted by the model usingkinetic parameters obtained in this work.

A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280 273

reaction still takes place on the grain surface, but the reaction rateis now limited by oxygen diffusion through the product layer. Athird step, based on the reaction of FeAl2O4 with H2 or CO in ahighly reducing atmosphere to form metallic iron, was not

considered for the reacting model since this step must not takeplace in a CLC process if complete combustion of the fuel gas isdesired, because of thermodynamic restrictions for H2 or CO con-version when FeAl2O4 is reduced to Fe0 [9].

ln (Ci)

-1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5

ln (

1/τ ch

)

-4.0

-3.5

-3.0

-2.5

-2.0

CH4

H2

CO

O2

nCH4= 0.25

nH2= 0.3

nCO= 0.6

nO2 = 0.9

Fig. 9. Plot of ln (Ci) vs ln (1/sch) to obtain the reaction order with respect to CH4, H2,CO and O2.

274 A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280

Taking the above comments and conditions in the TGA into con-sideration, the dependence of solids conversion with time for thisFe-based oxygen carrier is described by the following equation[55]:

t ¼ sch � XS;red þ splð1� 3 ð1� XS;redÞ2=3 þ 2 ð1� XS;redÞÞ ð7Þ

where sch and spl are the times for complete reduction conver-sion of the particle owing to the chemical reaction and the diffu-sion through the product layer. sch is calculated as follows:

sch ¼1

ks � Cni

ð8Þ

where ks is the chemical kinetic constant, which follows anArrhenius-type expression with temperature:

ks ¼ ks;0 e�EchRg �T ð9Þ

The time of complete conversion for the diffusion through theproduct layer reaction, spl, is defined by expression (10):

spl ¼1

Dpl � Cn0i

ð10Þ

In this experimental work, a sharp decrease in the reaction rateover time was observed when the diffusion through the productlayer of FeAl2O4 controlled the reduction reaction. This suggeststhat the effective product layer diffusivity, Dpl, is affected both bytemperature and solids conversion. In this case, an additional termto the effect with temperature is included, which modifies theeffective diffusivity through the product layer [56,57], with Dpl cal-culated as:

Dpl ¼ Dpl;0 e�kX �Xs;red e�EplRg T ð11Þ

where

kX ¼ kX;0 e�EXRg �T ð12Þ

With regard to oxidation kinetics, the reacting model assumes areaction rate only controlled by chemical reaction, since the reac-tion rate was very fast for the complete conversion of the oxygencarrier. The dependence of oxidation conversion on time for thisFe20cAl oxygen carrier is described by the following equation:

t ¼ sch � XS;ox ð13Þ

3.5. Determination of kinetic parameters

The kinetics parameters for reduction with CH4, H2, CO and oxi-dation with O2 were calculated with the kinetic model presented inthe previous section.

In a first stage, the kinetic parameters corresponding to the casewhere the reaction rate is controlled by the chemical reaction inthe gas–solid interphase were determined. Thus, the first step ofthe reaction is described by the following equation:

Fig. 8. Scheme of the reacting mechanism for CH

t ¼ sch � XS ð14Þ

sch values are obtained at different Ci by fitting each experimen-tal curve. Thus, the following equation was deduced from expres-sion (8).

ln1sch

� �¼ ln ks þ n � ln Ci ð15Þ

Fig. 9 shows a plot of ln(Ci) vs ln (1/sch) in order to calculate thereaction order, n, with respect to each reducing or oxidizing agent.This parameter was obtained for each gas from the slope of eachcurve. Furthermore, the chemical kinetic constant, ks, could beobtained from the intercept. The reaction order calculated forreduction with CH4 and oxidation with O2 were the lowest andthe highest, respectively. The reaction order for reduction withCH4 was very similar to that obtained by Mogthaderi and Song[34] with a Fe-based oxygen carrier supported on Al2O3 and pre-pared by mechanical mixing using the shrinking core model(SCM) for kinetic determination. In the case of reduction with H2

and CO, these authors found values that were considerably higher.With regard to oxidation reaction, Abad et al. [29] determined areaction order of 1 with O2 using a Fe-based oxygen carrier pre-pared by freeze granulation, which is a value very similar to theone found in this work.

By fitting the conversion vs time curves in Figs. 5 and 7 with themodel, parameter ks was determined at each temperature. TheArrhenius plot obtained from the reduction and oxidation reac-tions is shown in Fig. 10(a). From the slope of the curves, the acti-vation energy, Ech, was determined for each reaction. The highestvalue was obtained for CH4, i.e., 66 kJ/mol, whereas the activationenergies for H2, CO and O2 were considerably lower, ranging from 8

4 reduction with the Fe20cAl oxygen carrier.

A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280 275

to 23 kJ/mol. The value of the activation energy for oxidation reac-tion, i.e., 23 kJ/mol, was higher in comparison with other kineticstudies conducted in the literature on Fe-based oxygen carriers[21,29,36]. The opposite behavior was found with the activationenergies for reduction reactions with CO and H2. In these cases,the values determined in this work were slightly lower than theones found in the literature, see Table 1.

As mentioned previously, diffusion through the product layerbecame the limiting step that controlled the reduction reactionsat higher solids conversion values. The corresponding kineticparameters fitting the conversion curves for different gas concen-trations and temperatures were also calculated.

It was found that the reacting gas concentration had no influ-ence over the diffusion mechanism. Furthermore, it was observedthat the decay constant, kX, was not dependent on the reactingtemperature since there were no perceptible variations in the slopeof the conversion vs time curves when the reaction rate was con-trolled by the diffusion mechanism. It should be noted that param-eter kX affects the intensity of the reaction rate decrease in thesecond step, which was similar at all temperatures tested in thiswork. Therefore, the activation energy for kX was considered tobe EX = 0 J mol�1.

Fig. 10(b) illustrates a plot of ln (Dpl) vs the inverse of the tem-perature in order to determine the pre-exponential factor, Dpl,0, andthe activation energy, Epl, for the reduction reactions. Both param-eters achieved the highest values for CH4 reduction, see Table 3.

1/T (K-1) x 104

7.5 8.0 8.5 9.0 9.5

ln (

k s)

-3.75

-3.50

-3.25

-3.00

-2.75

-2.50

-2.25CH4

H2

CO

O2

(a)

ECH4 = 66 kJ/mol

EH2 = 8 kJ/mol

ECO = 14 kJ/mol

EO2 = 23 kJ/mol

1/T (K-1) x 104

7.5 8.0 8.5 9.0 9.5

ln (

Dp

l)

-6

-4

-2

0

2

4

6

8

10

(b) CH4

H2

CO

ECO = 164 kJ mol-1

ECH4 = 672 kJ mol-1

EH2 = 288 kJ mol-1

Fig. 10. Arrhenius plots to determine activation energy for ks (a), and effectiveproduct layer diffusivity, Dpl (b), for the reaction of CH4, H2, CO and O2 with theFe20cAl oxygen carrier.

Finally, the decay constant, kX, was also calculated. This parametertook the highest value for the reduction reaction with CH4, sincethe decrease in the reaction rate with time was slightly morenoticeable with this gas than with the other reducing gases. Table 3shows the kinetic parameters obtained in this work for reductionof Fe20cAl material with CH4, H2 and CO, as well as for oxidationwith O2.

As can be observed in Figs. 4–7, the theoretical curves deter-mined from the reaction model correctly predicted the experimen-tal results obtained during the whole reacting time for each type ofgas, temperature and concentration considered.

4. Discussion

The reaction kinetics and oxygen transport capacity determinedin this work can be used to evaluate relevant parameters for thedesign of a CLC system, such as the solids circulation rate betweenthe fuel reactor and the air reactor and the solids inventory in bothreactors.

The solids circulation rate must be high enough to transferoxygen for fuel combustion and to provide the heat necessary tomaintain optimum temperatures in the system. In the case of anFe-based oxygen carrier supported on Al2O3 for which the Fe2O3/Al2O3–FeAl2O4 redox couple is considered, heat plays a less impor-tant role because both reduction and oxidation reactions areexothermic [9]. Thus, there are no restrictions on the solids circu-lation rate in order to maintain a low temperature differencebetween both reactors [29].

In a previous work, Abad et al. [29] developed a simplifiedmodel to determine the solids inventory and the solids circulationrate in a CLC system taking into consideration the reactivity andoxygen transport capacity of the oxygen carrier particles. Thismethod enables the magnitude order of the design parametersto be established and, therefore, a comparison between oxygencarriers can be made by analyzing the obtained values.

4.1. Solids recirculation rate

The solids circulation rate depends on the oxygen carrier andthe fuel used, as well as on the variation of solids conversion inthe fuel reactor and air reactor. As previously mentioned, the recir-culation rate was calculated according to the model developed byAbad et al. [29]. This parameter is defined by Eq. (16), expressed asthe mass flow of fully oxidized oxygen carrier material, taking as areference the 1 MWth of power provided by the fuel and assumingcomplete gas conversion (DXg = 1).

_mOC ¼_mc

DXsð16Þ

where _mc is the characteristic circulation rate, a specific param-eter for each oxygen carrier–fuel combination.

_mc ¼2 � d �MO

ROC � DH0C

ð17Þ

The oxygen transport capacity of the material, ROC , is defined asROC ¼ RO � xMeO, where xMeO is the mass percentage of metal oxide inthe oxygen carrier. The active Fe2O3 content in the oxygen carrierwas 20 wt.% and RO ¼ 10 % for the Fe2O3–FeAl2O4 redox couple.Thus, the oxygen transport capacity was ROC = 0.02. The _mc param-eter took values of 4.0, 3.3 and 2.8 kg s�1 MW�1 for this materialwhen considering CH4, H2 and CO as fuels, respectively.

Fig. 11 illustrates the circulation rate, _mOC , as a function of thevariation of solids conversion in the fuel reactor if CH4 is used forreduction reaction. The solids circulation rate in a CLC unit is oftenevaluated by using the oxygen carrier to fuel ratio, /, which

Table 3Kinetic parameters for reaction of Fe20cAl oxygen carrier particles with reducing (CH4, H2, CO) and oxidizing gases (O2).

Units CH4 H2 CO O2

n Order of the reaction – 0.25 0.3 0.6 0.9ks,0 Pre-exponential factor of ks m3n mol�n s�1 4.34 � 101 1.45 � 10�1 1.59 � 10�1 3.64 � 10�1

Ech Activation energy for ks kJ/mol 66 8 14 23n0 Order of diffusion – 0 0 0 –Dpl,0 Pre-exponential factor of Dpl m3n0 mol�n0 s�1 9.80 � 1030 1.40 � 1013 2.29 � 109 –Epl Activation energy for Dpl kJ/mol 672 288 204 –kX,0 Pre-exponential factor of kX – 20 14 10 –EX Activation energy for kX kJ/mol 0 0 0 –

276 A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280

expresses the molar ratio between the potential flow oxygen in cir-culating oxygen carrier and the required flow of oxygen for fuelcombustion. Thus, the / parameter is also included in Fig. 11. Itcan be observed that _mOC parameter decreased as the variation ofsolids conversion, DXs, increased. The minimum circulation ratefor CH4 combustion with the Fe20cAl oxygen carrier at DXs = 1was 4.0 kg s�1 MW�1. On the other hand, the solids circulation flowwas dependent on the operating conditions and the configurationof the riser in the CLC unit. For usual conditions of velocity, tem-perature and excess of air, the maximum circulation rate in a CLCsystem is established at 16 kg s�1 MW�1 [29]. Therefore, a solidsconversion higher than 0.25 would be necessary for this materialin order to maintain the flow of solids under the average maximumlimit of 16 kg s�1 MW�1 if CH4 is used as the feed gas for the CLCunit.

Furthermore, it must be taken into consideration that tempera-ture greatly affected the reduction conversion degree of the oxygencarrier, and, consequently, the solids circulation rate. It can beinferred from Fig. 5 that, depending on the operating temperaturein the CLC system, the maximum variation of reduction conversionreached by this material in the fuel reactor would be different. As afirst approximation, it was assumed that the oxygen carrier wouldonly be reduced in a real CLC system up to solids conversion valueswithin the range at which the reaction kinetics was controlled bychemical reaction. Therefore, the maximum variation of reductionconversion was determined from experimental results for the dif-ferent reacting gases in a usual range of operating temperatures,see Fig. 12. From the results shown in Figs. 11 and 12, importantconclusions regarding adequate operating temperatures in the fuel

ΔXs

0.0 0.2 0.4 0.6 0.8 1.0

mo

c (k

g s

-1 M

W-1

)

0

10

20

30

40

1000 ºC

825 ºC16

5 2.5 1.67 1.25

φ

1273 K

1100 K

Fig. 11. Solids circulation rate vs variation of solids conversion in the fuel reactorusing the Fe20cAl material as oxygen carrier. Reducing gas: CH4.

reactor and solids circulation rates in the CLC system can bededuced if the Fe20cAl material is used as an oxygen carrier. ForCH4 combustion, it was concluded that DXs < 0.25 would not beallowed if a maximum limit of 16 kg s�1 MW�1 were assumed[29]. Therefore, temperatures lower than 1100 K could not be usedin the fuel reactor since the maximum variation of reduction con-version reached at this temperature (see Fig. 12) would require asolids circulation rate higher than the upper limit. On the contrary,at temperatures higher than 1273 K, the corresponding values ofDXS,ch and solids circulation rate were 0.57 and 7.0 kg s�1 MW�1

respectively.

4.2. Solids inventory

In a CLC process, it is desirable to minimize the amount of oxy-gen carrier existing in the fuel and air reactors in order to reducethe size of the plant fan power and the cost relative to the oxygencarrier. The solids inventory can be determined from a mass bal-ance to the oxygen carrier and fuel gas in both reactors in combi-nation with reaction kinetics. Furthermore, the solids inventory isalso directly related to the reactivity and oxygen transport capacityof the oxygen carrier.

As in the case of the determination of the solids circulation rate,the calculation of the solids inventory was also based on the meth-odology developed by Abad et al. [29]. The solids inventory in thefuel reactor and air reactor, mOC;FR and mOC;AR, can be expressedaccording to Eqs. (18) and (19):

mOC;FR ¼ _mcsr

UFRð18Þ

Temperature (K)

1050 1100 1150 1200 1250 1300

ΔXs,

red

ch

0.0

0.2

0.4

0.6

0.8

1.0

CH4

H2

CO

Fig. 12. Variation of reduction conversion in chemically controlled reaction forFe20cAl material reached in the fuel reactor for CH4, H2 and CO combustion in ausual range of operating temperatures for CLC.

Table 4Minimum solids inventories for CH4, H2 and CO combustion in fuel and air reactors ofa CLC unit with Fe20cAl as oxygen carrier (T = 1223 K).

CH4 H2 CO

Min. solids inventory in FR (kg) 54 37 42Min. solids inventory in AR (kg) 95 79 68Min. total solids inventory (kg) 149 116 109

A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280 277

mOC;AR ¼ _mcso

UARð19Þ

where sr and so are the times for complete conversion of theparticles in the fuel reactor and air reactor, respectively. Theseparameters are obtained at an average reacting gas concentrationcalculated from the following expression:

�Ci ¼DXg � Cn

i;0R Xg;outXg;in

1þeg �Xg

1�Xg

h indXg

ð20Þ

where eg represents the expansion of volume as a result of thechemical reaction and can be calculated as:

eg ¼Vg;Xg¼1 � Vg;Xg¼0

Vg;Xg¼0ð21Þ

where Vg;Xg¼0 and Vg;Xg¼1 are the volumes of the gas mixture at Xg = 0and Xg = 1, respectively. This parameter changes its value dependingon the fuel gas considered for the CLC process. For example, eg takesa value of 2 for CH4, 0 for H2 and CO and �0.21 for the oxidationreaction with air.

The average concentrations were determined at 1223 K in bothreactors taking into consideration the reaction order, n, shown inTable 3. When 100 vol.% fuel gas concentration at the fuel reactorinlet and a final gas conversion of 0.999 was considered, the aver-age CH4, H2 and CO concentrations in this reactor were 15.7, 30.7and 24.6 vol.%, respectively. On other hand, for an oxidation reac-tion in the air reactor with 21 vol.% O2 and 20 vol.% excess air,the resulting average O2 concentration was 11.2 vol.%. The timesfor complete conversion of the particles in the fuel reactor, consid-ering a temperature of 1223 K, were 13.7, 11.0 and 14.2 s for CH4,H2 and CO combustion, respectively. In the case of oxidationreaction, the time for complete conversion of the Fe20cAl oxygencarrier particles in the air reactor at 1223 K was 23.9 s.

Eqs. (18) and (19) are also dependent on parameters UFR and UAR

which are the characteristic reactivity in the fuel reactor and airreactor respectively. Assuming perfect mixing of the solids in bothreactors, Uj can be expressed as a function of the variation of solidsconversion and as a function of the solids conversion in the inlet ofeach reactor. For the kinetic model used in this work, the expressionsfor Uj are the following and take a value between 0 and 1:

UFR ¼ 1� exp �ð1� XS;red in FRÞDXS

UFR

� �ð22Þ

UAR ¼ 1� exp �ð1� XS;ox in ARÞDXS

UAR

� �ð23Þ

The total solids inventory in the CLC system, defined in terms ofkilograms of fully oxidized oxygen carrier per MW of fuel, can becalculated adding the masses obtained in the fuel reactor and airreactor:

mOC ¼ mOC;FR þmOC;AR ð24Þ

For comparison purposes, the minimum solids inventory in theCLC system was calculated for the three reducing gases consideredin this work, i.e., CH4, H2 and CO considering fuel and air reactors at1223 K. In this case, DXS,red ch = 0.52, and the active Fe2O3 contentwas 10.4 wt.%; see Fig. 12. Thus, ROC = 1.04 wt.%. But the reactingtime to achieve the solids conversion DXS,red ch, i.e. s0r to be usedin Eq. (18) instead of sr , was conveniently reduced by (seeAppendix A)

s0r ¼ DXs;red chsr ð25Þ

The minimum solids inventory was obtained when the varia-tion of solids conversion in the reactors was very low (DXS ? 0)and the characteristic reactivity reached the highest value. For

the kinetic model used in this work, the highest value of parameterUj is 1. The minimum solids inventories for CH4, H2 and CO com-bustion were 149, 116 and 109 kg MW�1, respectively, see Table 4.A comparison between the expected performance of different oxy-gen carriers can easily be made by contrasting the minimum solidsinventory values. The minimum solids inventories obtained for theFe20cAl material were compared to the ones found for the Fe45Aloxygen carrier prepared by freeze granulation [29] or ilmenite [37].In fact, the high reactivity of Fe20cAl with CH4 resulted in a lowersolids inventory in the fuel reactor for Fe20cAl (54 kg/MWth) com-pared to that of Fe45Al (950 kg/MWth) and of activated ilmenite(272 kg/MWth), even when the iron content in these materialswas higher. Thus, it can be concluded that the Fe20cAl oxygen car-rier, prepared by the incipient impregnation method, presentsmuch higher reactivity with CH4, and hence the minimum solidsinventory needed in the CLC unit reactor is lower in comparisonwith the other Fe-based material.

In real conditions, the solids inventory depends on the degree ofoxidation of the oxygen carrier particles in the inlet of both reac-tors and on the value of DXS. Furthermore, it must be also takeninto consideration that in real conditions, the variation of solidsconversion, DXS, is defined by the solids circulation rate estab-lished in the CLC system, see Eq. (16), and it is possible that theoxygen carrier will not completely oxidized when it enters the fuelreactor .

Fig. 13 illustrates the total solids inventory of the Fe20cAl oxy-gen carrier needed in a CLC unit for CH4 combustion as a functionof the oxidation conversion of the particles at the inlet of the fuelreactor considering that the variation of solids conversion, DXS, islimited to the minimum value related to the maximum solidscirculation rate of 16 kg s�1 MW�1. Thus, these calculations weremade for a variation of solids conversion of 0.25, see Fig. 11. Thisvalue of DXS is adequate because it is within the optimum rangeproposed by Abad et al. [29], i.e., DXS = 0.2–0.5, to obtain low solidsinventories and reasonable circulation rate values.

Now, the method to calculate UFR was modified from thatshown by Abad et al. [29] to consider the active Fe2O3 fraction asbeing only the fraction reacting with chemical control, but thesolid fraction reacting with slow diffusional controlled reactionwas assumed to be inert; see Appendix A. Thus, to calculate UFR

by Eq. (22), a modified conversion X0S; red was used, which rangesfrom 0 to 1 when the solids conversion varies from 0 to XS; red ch.Considering the case of CH4 combustion at 1223 K, XS;red ch wasestablished at 0.52 approximately (see Fig. 12), which means thatfor a variation of solids conversion of DXS ¼ 0:25, DX0S takes a valueof 0.48.

With the above considerations, the inlet conversion of solids tothe fuel reactor must fulfill XS;ox in FR > DXS þ ð1� XS;red chÞ becausethe fraction ð1� XS;red chÞ is considered inert for the reduction reac-tion. It can be observed in Fig. 13 that the minimum oxidation con-version at the inlet of the fuel reactor is XS;ox in FR ¼ 0:73 forDXS ¼ 0:25. If the value of parameter XS;ox in FR is very close to theaforementioned limit, the availability of oxygen for the reductionconversion will be low because the oxygen carrier particles arereduced to a great extent and, consequently, the required solidsinventory in this reactor will tend towards infinity. At the same

Xs,ox in FR

0.70 0.75 0.80 0.85 0.90 0.95 1.00

So

lids

inve

nto

ry (

kg c

arri

er M

W-1

)

0

100

200

300

400

500

FR

ARTotal

Fig. 13. Total solids inventory of the Fe20cAl material as a function of the solidsconversion at the inlet of the fuel reactor (XS;ox in;FR). Reducing agent: CH4. DXS = 0.25.T = 1223 K. It is assumed that the Fe20cAl oxygen carrier is only reduced up tosolids conversion values within the range at which the reaction kinetics iscontrolled by chemical reaction.

278 A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280

time, a low value of XS;ox in FR implies that the solids inventoryneeded in the air reactor decreases. The opposite reasoning canbe made for very high values of oxidation conversion at the fuelreactor inlet. The combination of both effects over the total solidsinventory entails the presence of a minimum which correspondsto the optimum solids inventory in the CLC system under theseconditions. In this case, the minimum amount of this material is272 kg MW�1 at an approximated value of XS;ox in FR ¼ 0:85.

Finally, it should be remarked that the value of solids inventorycalculated for a real CLC system would be from 2 to 10 times higher[58] because the method for determination does not take intoconsideration such fluidization effects as the resistance to gasexchange between the bubbles and the emulsion in the fluidizedbeds. For example, Gayán et al. [16] found that for this oxygencarrier a solids inventory of 500 kg MW�1 was necessary for com-plete combustion of CH4 in the fuel reactor of a 500 Wth continuousCLC unit. Thus, the total solids inventory in this CLC unit would bewithin the range established by Abad et al. [58] taking into consid-eration the value calculated in this work. Therefore, these valuesare valid for comparison purposes with other oxygen carriers, butnot for design purposes.

The results obtained from this simplified model provide valu-able information for comparison purposes with different materialsand a first approximation to the total solids inventory needed in areal CLC system. For example, when using this calculation method,the Fe20cAl oxygen carrier presents a much lower value of totalsolids inventory for CH4 combustion in comparison with anotherFe-based oxygen carrier supported on Al2O3 with 60 wt.% Fe2O3

prepared by freeze granulation [29]. This fact emphasizes the highreactivity achieved by oxygen carriers prepared by means of theimpregnation method. On other hand, the solids inventoryobtained for the Fe20cAl material is higher than the ones foundfor Ni- or Cu-based oxygen carriers in that work [29]. However,it must be considered that iron is significantly cheaper and moreenvironmentally friendly than either nickel or copper.

5. Conclusions

The isothermal method in a TGA was identified as the propermethod with which to obtain the reduction and oxidation kineticsof the Fe20cAl oxygen carrier with CH4, H2, CO and O2 as reacting

gases under typical CLC conditions. The effect of temperature andthe concentration of the reacting gases on reaction rate of Fe20cAlmaterial were studied.

The Fe20cAl oxygen carrier was always reduced to FeAl2O4 attypical temperatures in CLC regardless of the final solids conver-sion achieved in the oxygen carrier samples. The formation ofFe3O4 as a reduced Fe-based phase was never found in the parti-cles. This result suggested that the reduction reaction mechanismwas based on the interaction of Fe2O3 with Al2O3 in presence ofthe reacting gases to form FeAl2O4 as the only stable Fe-basedphase. Furthermore, it was found that the reduction rate and thefinal solids conversion were clearly affected by the temperature.Nevertheless, complete regeneration was found after the oxidationstep with O2.

The grain model with uniform conversion in the particle andreaction in grains following the shrinking core model (SCM) wasused for kinetics determination, assuming a first step of reductioncontrolled by chemical reaction, followed by a second step con-trolled by the diffusion through the product layer around thegrains. The reaction order values found for the reducing gases ran-ged between 0.25 and 0.6. Additionally, the lowest and the highestactivation energy values were found for reduction reactions withH2 and CH4 respectively, 8 and 66 kJ/mol. With regard to oxidationkinetics, the reacting model assumed a reaction rate that was onlycontrolled by chemical reaction. The reaction order took a value of0.9 and the activation energy for oxidation reaction with O2 was23 kJ/mol.

Finally, the solids inventory needed in a CLC system was alsoestimated taking into consideration the results obtained from thekinetics study. Firstly, it was found that the minimum solidsinventories for CH4, H2 and CO combustion with the Fe20cAloxygen carrier were 149, 116 and 109 kg MW�1, respectively. Fur-thermore, the solids inventory was calculated for usual operatingconditions in a CLC unit, such as DXs = 0.25 and a temperature inthe fuel reactor and air reactor of 1223 K. Under these conditions,and considering CH4 as fuel gas, the minimum total solids inven-tory was 272 kg MW�1.

Acknowledgements

This paper is based on the work carried out within the frame-work of the SUCCESS project, funded by the European Commissionunder the seventh Framework Programme (Contract 608571). Thisresearch was supported by the Spanish Ministry of Science andInnovation (MICINN Project: ENE2011-26354) and by FEDER.

Appendix A. Determination of solids inventory. Modification tothe calculation method proposed by Abad et al. [29] consideringthat complete reduction of the oxygen carrier was not possibleowing to diffusional limitations

Considering the reaction kinetics obtained in this work, it canbe assumed that diffusion-controlled reaction does not take placeto any relevant extension in a CLC unit because this takes muchmore time than the average residence time of particles in the fuelreactor. Thus, additional expressions were defined in order to cal-culate the total solids inventory in the fuel reactor for the casewhere the oxygen carrier is only reduced up to solids conversionvalues within the range at which the reduction is controlled bychemical reaction.

Thus, the solids conversion had to be recalculated to be used inEqs. (22) and (23). Namely, the active Fe2O3 content, and thereforethe actual oxygen transport capacity R0OC , depends on the XS,red ch

value shown in Fig. 12 as

R0OC ¼ XS;red ch � ROC ðA:1Þ

t

Xs,red

1

Xs,red ch

τrτ'r

(a)

t

X's,red

1

τ'r

(b)

Fig. A1. (a) Determination of s0r and XS;red ch parameters from reduction conversion vs time curve (Fig. 5a). (b) Normalized reduction conversion vs time curve when thereduction reaction is controlled by chemical reaction. Operating conditions: reducing gas, 15 vol.% CH4; temperature: 1223 K.

A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280 279

Therefore, the solid fraction reacting with slow diffusion-controlled reaction is assumed to be inert. The solids inventoryin the fuel reactor can be calculated in these conditions as follows:

mOC;FR ¼2 � d �MO

DH0C

� s0rR0OC

� 1U0FR

ðA:2Þ

Fig. A1(a) illustrates an example of a conversion vs time curvefor CH4 combustion from which both s0r and XS;red ch parameterscan be calculated. XS;red ch is the maximum conversion reached bythe oxygen carrier particles when the reduction reaction iscontrolled by chemical reaction. The time corresponding to thisreduction conversion is defined as s0r . The calculation methodconsiders a modified conversion X 0S; red, which ranges from 0 to 1when the solids conversion varies from 0 to XS; red ch. s0r can be deter-mined at each temperature and reducing gas concentration withthe kinetic data obtained in this work as

s0r ¼ DXs;red chsr ðA:3Þ

Thus, it was deduced that

sr

ROC¼ s0r

R0OC

ðA:4Þ

and consequently, Eq. (A.2) can be written as

mOC;FR ¼2 � d �MO

DH0C

� sr

ROC� 1U0FR

ðA:5Þ

The characteristic reactivity in the fuel reactor was also rede-fined according to the following expression:

U0FR ¼ 1� exp �ð1� X0S;red in FRÞ

DX0SU0FR

!ðA:6Þ

where X 0S;red in FR and DX 0S are calculated as (see Fig. A1(b)):

X0S;red in FR ¼XS;red in FR

XS;red ch; XS;red in FR < XS;red ch ðA:7Þ

DX0S ¼DXS

XS;red ch; DXS < XS;red ch ðA:8Þ

Furthermore, X 0s;ox in FR can be calculated as:

X0S;ox in FR ¼XS;ox in FR � XS;ox min

1� XS;ox minðA:9Þ

Considering that XS;ox min ¼ 1� XS;red ch, XS;ox in FR can be definedaccording to Eq. (A.10) as follows:

XS;ox in FR ¼ X0S;ox in FR � XS;red ch þ ð1� XS;red chÞ ðA:10Þ

References

[1] IPCC, 2013. Climate Change 2013: The Physical Science Basis. TechnicalSummary. Working Group I Contribution to the IPCC Fifth Assessment Report(AR5).

[2] IPCC, 2005, in: B. Metz, O. Davidson, H. de Coninck, M. Loos, L. Meyer (Eds.),Special report on carbon dioxide capture and storage. Working group II of theIntergovernmental Panel on Climate Change, Cambridge University Press,Cambridge, UK and New York, NY, USA, 2005.

[3] W.K. Lewis, E.R. Gilliland, Production of pure carbon dioxide, S.O.D Company,US Patent: 2 665 971, United States, 1954.

[4] H.J. Richter, K. Knoche, Reversibility of combustion processes. Efficiency andcosting – second law analysis of processes, ACS Symp. Ser. 235 (1983) 71–85.

[5] M. Ishida, D. Zheng, T. Akehata, Evaluation of a chemical-looping-combustionpower-generation system by graphic exergy analysis, Energy 12 (1987) 147–154.

[6] M. Ishida, H. Jin, A new advanced power-generation system using chemical-looping combustion, Energy 19 (1994) 415–422.

[7] M. Ishida, H. Jin, A novel combustor based on chemical-looping reactions andits reaction kinetics, J. Chem. Eng. Jpn. 27 (1994) 296–301.

[8] A. Lyngfelt, B. Leckner, T. Mattisson, A fluidized-bed combustion process withinherent CO2 separation; application of chemical-looping combustion, Chem.Eng. Sci. 56 (2001) 3013–3101.

[9] J. Adánez, A. Abad, F. García-Labiano, P. Gayán, L.F. de Diego, Progress inchemical looping combustion and reforming technologies, Prog. EnergyCombust. Sci. 38 (2012) 215–282.

[10] J. Adánez, L.F. de Diego, F. García-Labiano, P. Gayán, A. Abad, J.M. Palacios,Selection of oxygen carriers for chemical-looping combustion, Energy Fuels 18(2004) 371–377.

[11] T. Mattisson, M. Johansson, A. Lyngfelt, Multicycle reduction and oxidation ofdifferent types of iron oxide particles – application to chemical-loopingcombustion, Energy Fuels 18 (2004) 628–637.

[12] M. Johansson, T. Mattisson, A. Lyngfelt, Investigation of Fe2O3 with MgAl2O4

for chemical-looping combustion, Ind. Eng. Chem. Res. 43 (2004) 6978–6987.[13] F. García-Labiano, J. Adánez, L.F. de Diego, P. Gayán, A. Abad, Effect of pressure

on the behaviour of copper-, iron-, and nickel-based oxygen carriers forchemical looping combustion, Energy Fuels 20 (2006) 26–33.

[14] A. Abad, F. García-Labiano, L.F. de Diego, P. Gayán, J. Adánez, Reduction kineticsof Cu-, Ni- and Fe-based oxygen carriers using syngas (CO+H2) for chemicallooping combustion, Energy Fuels 21 (2007) 1843–1853.

[15] M. Ortiz, L.F. de Diego, P. Gayán, M.A. Pans, F. García-Labiano, A. Abad, J.Adánez, Hydrogen production coupled with CO2 capture by chemical-loopingusing mixed Fe-Ni oxygen carriers, in: Proc 1st Int Conf on Chemical Looping,Lyon, France, 2010.

[16] P. Gayán, M.A. Pans, M. Ortiz, A. Abad, L.F. de Diego, F. García-Labiano, J.Adánez, Testing of a highly reactive impregnated Fe2O3/Al2O3 oxygen carrierfor a SR-CLC system in a continuous CLC unit, Fuel Process. Technol. 96 (2012)37–47.

[17] A. Cabello, C. Dueso, F. García-Labiano, P. Gayán, A. Abad, L.F. de Diego, J.Adánez, Performance of a highly reactive impregnated Fe2O3/Al2O3 oxygencarrier with CH4 and H2S in a 500 Wth CLC unit, Fuel 121 (2014) 117–125.

[18] P. Cho, T. Mattisson, A. Lyngfelt, Carbon formation on nickel and iron oxide-containing oxygen carriers for chemical-looping combustion, Ind. Eng. Chem.Res. 44 (2005) 668–676.

[19] L.F. de Diego, F. García-Labiano, J. Adánez, P. Gayán, A. Abad, G. Sprachmann,Performance of Cu- and Fe-based oxygen carriers in a 500 Wth CLC unit forsour gas combustion with high H2S content, Int. J. Greenhouse Gas Control 28(2014) 168–179.

[20] F. García-Labiano, J. Adánez, L.F. de Diego, P. Gayán, A. Abad, A. Cabello, G.Sprachmann, Energy exploitation of acid gas with high H2S content by meansof a chemical looping combustion system, submitted for publication.

280 A. Cabello et al. / Chemical Engineering Journal 258 (2014) 265–280

[21] S.R. Son, S.D. Kim, Chemical-looping combustion with NiO and Fe2O3 in athermobalance and circulating fluidized bed reactor with double loops, Ind.Eng. Chem. Res. 45 (2006) 2689–2696.

[22] A. Abad, T. Mattisson, A. Lyngfelt, M. Johansson, The use of iron oxide asoxygen carrier in a chemical-looping reactor, Fuel 86 (2007) 1021–1035.

[23] A. Lyngfelt, H. Thunman, Construction and 100 h of operational experience of a10-kW chemical-looping combustor, in: D.C. Thomas, S.M. Benson (Eds.),Carbon dioxide capture for storage in deep geologic formations - Results fromthe CO2 capture project, vol. 1, Elsevier, Oxford, UK, 2005 (chapter 36).

[24] T. Pröll, K. Mayer, J. Bolhàr-Nordenkampf, P. Kolbitsch, T. Mattisson, A.Lyngfelt, H. Hofbauer, Natural minerals as oxygen carriers for chemical loopingcombustion in a dual circulating fluidized bed system, Energy Proc. 1 (2009)27–34.

[25] P. Kolbitsch, J. Bolhàr-Nordenkampf, T. Pröll, H. Hofbauer, Operatingexperience with chemical looping combustion in a 120 kW dual circulatingfluidized bed (DCFB) unit, Int. J. Greenhouse Gas Control 4 (2010) 180–185.

[26] M. Rydén, M. Johansson, E. Cleverstam, A. Lyngfelt, T. Mattisson, Ilmenite withaddition of NiO as oxygen carrier for chemical-looping combustion, Fuel 89(2010) 3523–3533.

[27] P. Moldenhauer, M. Rydén, A. Lyngfelt, Testing of minerals and industrial by-products as oxygen carriers for chemical-looping combustion in a circulatingfluidized-bed 300 W laboratory reactor, Fuel 93 (2012) 351–363.

[28] M. Ortiz, P. Gayán, L.F. de Diego, F. García-Labiano, A. Abad, M.A. Pans, J.Adánez, Hydrogen production with CO2 capture by coupling steam reformingof methane and chemical-looping combustion: Use of an iron-based wasteproduct as oxygen carrier burning a PSA tail gas, J. Power Sources 196 (2011)4370–4381.

[29] A. Abad, J. Adánez, F. García-Labiano, L.F. de Diego, P. Gayán, J. Celaya, Mappingof the range of operational conditions for Cu-, Fe-, and Ni-based oxygencarriers in chemical-looping combustion, Chem. Eng. Sci. 62 (2007) 533–549.

[30] K.S. Go, S.R. Son, S.D. Kim, Reaction kinetics of reduction and oxidation ofmetal oxides for hydrogen production, J Hydrogen Energy 33 (2008) 5986–5995.

[31] C.D. Bohn, J.P. Cleeton, C.M. Müller, J.F. Davidson, A.N. Hayhurst, S.A. Scott, J.S.Dennis, The kinetics of the reduction of iron oxide by carbon monoxide mixedwith carbon dioxide, AIChE J. 56 (2010) 1016–1029.

[32] J. Bao, Z. Li, H. Sun, N. Cai, Experiment and rate equation modeling of Feoxidation kinetics in chemical looping combustion, Combust. Flame 160(2013) 808–817.

[33] E.R. Monazam, R.W. Breault, R. Siriwardane, G. Richards, S. Carpenter, Kineticsof the reduction of hematite (Fe2O3) by methane (CH4) during chemicallooping combustion: A global mechanism, Chem. Eng. J. 232 (2013) 478–487.

[34] B. Moghtaderi, H. Song, Reduction properties of physically mixed metallicoxide oxygen carriers in chemical looping combustion, Energy Fuels 24 (2010)5359–5368.

[35] E.R. Monazam, R.W. Breault, R. Siriwardane, D.D. Miller, Thermogravimetricanalysis of modified hematite by methane (CH4) for chemical-loopingcombustion: a global kinetics mechanism, Ind. Eng. Chem. Res. 52 (2013)14808–14816.

[36] E. Ksepko, M. Sciazko, P. Babinski, Studies on the redox reaction kinetics ofFe2O3–CuO/Al2O3 and Fe2O3/TiO2 oxygen carriers, Appl. Energy 115 (2014)374–383.

[37] A. Abad, J. Adánez, A. Cuadrat, F. García-Labiano, P. Gayán, L.F. de Diego,Reaction kinetics of ilmenite for chemical-looping combustion, Chem. Eng. Sci.66 (2011) 689–702.

[38] S. Nasr, K.P. Plucknet, Kinetics of iron ore reduction by methane for chemicallooping combustion, Energy Fuels 28 (2014) 1387–1395.

[39] C. Dong, S. Sheng, W. Qin, Q. Lu, Y. Zhao, X. Wang, J. Zhang, Density functionaltheory study on activity of a-Fe2O3 in chemical-looping combustion system,Appl. Surf. Sci. 257 (2011) 8647–8652.

[40] F. Li, Z. Sun, S. Luo, L.S. Fan, Ionic diffusion in the oxidation of iron-effect ofsupport and its implications to chemical looping applications, Energy Environ.Sci. 4 (2011) 876–880.

[41] F. Li, S. Luo, Z. Sun, X. Bao, L.S. Fan, Role of metal oxide support in redoxreactions of iron oxide for chemical looping applications: experiments anddensity functional theory calculations, Energy Environ. Sci. 4 (2011) 3661–3667.

[42] C. Dong, X. Liu, W. Qin, Q. Lu, X. Wang, S. Shi, Y. Yang, Deep reduction behaviorof iron oxide and its effect on direct CO oxidation, Appl. Surf. Sci. 258 (2012)2562–2569.

[43] Q. Tan, W. Qin, Q. Chen, C. Dong, W. Li, Y. Yang, Synergetic effect of ZrO2 on theoxidation–reduction reaction of Fe2O3 during chemical looping combustion,Appl. Surf. Sci. 258 (2012) 10022–10027.

[44] W. Li, Q. Chen, T. Mi, Density functional theory study of interaction betweenCO and oxygen carrier Fe2O3@SBA-15, Adv. Mater. Res. 610 (2013) 498–501.

[45] D.D. Miller, R. Siriwardane, Mechanism of methane chemical loopingcombustion with hematite promoted with CeO2, Energy Fuels 27 (2013)4087–4096.

[46] W. Qin, Q. Chen, Y. Wang, C. Dong, J. Zhang, W. Li, Y. Yang, Theoretical study ofoxidation–reduction reaction of Fe2O3 supported on MgO during chemicallooping combustion, Appl. Surf. Sci. 266 (2013) 350–354.

[47] W. Qin, Y. Wang, C. Dong, J. Zhang, Q. Chen, Y. Yang, The synergetic effect ofmetal oxide support on Fe2O3 for chemical looping combustion: a theoreticalstudy, Appl. Surf. Sci. 282 (2013) 718–723.

[48] L. Wang, Q. Li, W. Qin, Z. Zheng, X. Xiao, C. Dong, Activity of Fe2O3 and its effecton CO oxidation in the chemical looping combustion: a theoretical account,Adv. Mater. Res. 726 (2013) 2040–2044.

[49] F. García-Labiano, L.F. de Diego, J. Adánez, A. Abad, P. Gayán, Temperaturevariations in the oxygen carrier particles during their reduction and oxidationin a chemical looping combustion system, Chem. Eng. Sci. 60 (2005) 851–862.

[50] HSC Chemistry 6.1, Chemical Reaction and Equilibrium Software withThermochemical Database and Simulation Module, Oututec Research Oy.,Pori, Finland, 2008.

[51] Q. Zafar, A. Abad, T. Mattisson, B. Gevert, Reaction kinetics of a freeze-granulated NiO/MgAl2O4 oxygen carrier particles for chemical-loopingcombustion, Energy Fuels 21 (2007) 610–618.

[52] F. García-Labiano, L.F. de Diego, J. Adánez, A. Abad, P. Gayán, Reduction andoxidation kinetics of a copper-based oxygen carrier prepared by impregnationfor chemical-looping combustion, Ind. Eng. Chem. Res. 43 (2004) 8168–8177.

[53] J.E. Readman, A. Olafsen, J.B. Smith, R. Blom, Chemical looping combustionusing NiO/NiAl2O4: mechanisms and kinetics of reduction–oxidation (Red–Ox)reactions from in situ powder X-ray diffraction and thermogravimetryexperiments, Energy Fuels 20 (2006) 1382–1387.

[54] C. Dueso, M. Ortiz, A. Abad, F. García-Labiano, L.F. de Diego, P. Gayán, J. Adánez,Reduction and oxidation kinetics of nickel-based oxygen-carriers for chemical-looping combustion and chemical-looping reforming, Chem. Eng. J. 188 (2012)142–154.

[55] O. Levenspiel, Chemical Reaction Engineering, John Wiley and Sons, New York,1981.

[56] L.F. de Diego, A. Abad, A. Cabello, P. Gayán, F. García-Labiano, J. Adánez,Reduction and oxidation kinetics of a CaMn0.9Mg0.1O3�d oxygen carrier forchemical-looping combustion, Ind. Eng. Chem. Res. 53 (2014) 87–103.

[57] J. Adánez, F. García-Labiano, A. Abad, L.F. de Diego, P. Gayán, Regeneration ofsulfided dolomite with steam and carbon dioxide, Energy Fuels 15 (2001) 85–94.

[58] A. Abad, J. Adánez, F. García-Labiano, L.F. de Diego, P. Gayán, Modelling of thechemical-looping combustion of methane using a Cu-based oxygen-carrier,Combust. Flame 3 (2010) 602–615.