(664d) CO2 Utilization Via Chemical Looping Process | AIChE

(664d) CO2 Utilization Via Chemical Looping Process

Authors 

Galvita, V. V. - Presenter, Ghent University
Poelman, H., Ghent University
Marin, G. B., Ghent University



Introduction

Chemical looping is one of several emerging
technology options capable to reduce the carbon dioxide emissions and helping
in a diverse range of applications for production of fuels, chemicals and
electricity [1-3]. In a general chemical
looping process, a metal oxide (MeO) is circulated between two reactors - the
reducer and oxidizer. For a combustion process, MeO reacts in the reducer with
the fuel to produce CO2, H2O and leaving metal (Me) or
metal oxide in reduced state [1]. In the
oxidizer it is re-oxidized to its initial state with air. Thus, fuel and air
never mix and CO2 is not diluted by nitrogen.

Other chemical looping schemes can produce hydrogen
or synthesis gas. In chemical looping steam (CLSR) or dry reforming (CLDR)
processes the fuel is again completely oxidized over the metal oxides (Eq. 1). The
H2O or CO2 product is used as oxidant instead of air. The
outlet from the looping reactor then consists of H2 or CO.

A key issue in CLDR processes is the selection of an
appropriate oxygen carrier material. The reactivity of the oxygen storage
material, its cost and thermal stability are critical selection criteria. A
broad range of metals was analyzed to determine the thermodynamic equilibrium
limitations for redox cyclic processes [4-7].
While a number of metals/metal oxides (Mo, Cr, Zn, Co, Nb, Ce) give reasonable
CO2 reduction capacity, it is apparent that only iron yields high
oxygen storage capacity from CO2 (0.7 mol CO2/ mol Fe)
over a wide range of operating temperatures (600-1800°C).

Thus, the chemical looping conversion of CO2
into CO or CO2 and H2O mixture to syngas over iron oxide
as oxygen storage material has been one of the possible target technologies for
CO2 utilization. The ability of the oxygen storage material to
maintain its high activity in repeated reduction/re-oxidation cycles is the
most critical issue for the overall economic viability of the CO2
splitting process.

The objective of the present study is to investigate
the CO2 re-oxidation efficiency of a series of CeO2-Fe2O3
mixed oxides prepared by co-precipitation for the utilization of CO2
in a chemical looping process. Experimental performance data are reported as
well as structural characterization of the oxygen storage materials during
reduction and CO2 re-oxidation. In addition, the stability of
performance was tested through prolonged use of the material.

Results

Mixed
CeO2-Fe2O3 materials with 0, 20, 50,70, 90 and
100 wt% CeO2 were prepared for CO2 transformation to CO
by means of chemical looping. The samples were characterized with HRTEM, EDX,
SAED, EEL and in situ XRD and tested in cycled reactions for activity and
stability. The generation of CO from CO2 through H2-CO2
redox cycles was investigated at 600oC to evaluate the effect of CeO2
upon Fe2O3. Two different phases were observed in the CeO2-Fe2O3
samples. For a CeO2 content below 70wt%, a distinct Fe2O3
phase was always present. The second phase appeared in all mixed CeO2-Fe2O3
samples and was identified as a nano-crystalline solid solution of iron in
ceria with 20 nm crystallite size and an atomic ratio of Fe:Ce up to 1:5. In
all mixed CeO2-Fe2O3 samples, the solid
solution crystallite size was considerably smaller compared to pure ceria, while
a size decrease for Fe2O3 was noticed with increasing CeO2
loading. The crystallographic structure of all CeO2-Fe2O3
was studied under H2 reduction and CO2 oxidation, using
time-resolved in situ X-ray diffraction. In a 90wt% CeO2 sample, no
iron related diffractions were present at the start, but the H2-TPR
treatment made Fe appear as phase segregating from ceria. For lower CeO2
loadings, H2-TPR reduced the separate Fe2O3
phase to Fe3O4, FeO and Fe. The solid solution phase was
always partially reduced. In CO2-TPO, re-oxidation of iron followed
a one-step pathway to Fe3O4 around 500oC,
while for the solid solution re-oxidation occurred at 450°C. The addition
of CeO2 to Fe2O3 had a beneficial effect upon
the activity and stability of the material. Reduction-re-oxidation cycles could
be repeated many times without significant loss of CO2 conversion
efficiency.

The activity and stability of the mixed CeO2-Fe2O3
samples as compared to pure Fe2O3 and CeO2 a
test reaction with H2 reduction and CO2 re-oxidation was
set up. The samples are reduced in 5%H2 in Ar during 1 min. By
contacting the reduced sample with CO2, CO is produced. As an
example, the observed space time yield of CO as a function of time for 20wt% CeO2-
Fe2O3during
the 2nd, 5th and 10th cycle is displayed in Fig.
1a.

AIChE_firure_1.jpg

Figure
1. (a) Space time yield of CO during the re-oxidation phase with CO2/He
for 20wt% CeO2-Fe2O3 after 1, 5 and 10 redox cycles; (b) CO
produced after 1st (dark bar) and 10th cycle (light bar) over
the xCe-Fe samples.

Fig. 1b shows the
amount of produced CO after the 1st and 10th redox cycle
for samples 0 to 100wt%CeO2-Fe2O3. In both cycles, the product yield is higher for mixed materials
than for pure ones. Especially the samples 20 to 70wt% CeO2-Fe2O3 produce over 5 times more CO than Fe2O3, with
the highest yield for 20 wt%CeO2-Fe2O3. The amount of CO produced over CeO2 was 10 times lower
than for Fe2O3. Highest CO yield
was obtained for 20wt% CeO2-Fe2O3, while
70 and 50wt% CeO2-Fe2O3 were the most
stable combinations. The addition of ceria to Fe2O3 thus
formed a solid solution phase, yielding improved activity. Moreover, the
combination of both oxides efficiently suppressed the sintering effect,
resulting in improved stability in chemical looping processes.

Acknowledgment

This work was supported by the ‘Long Term
Structural Methusalem Funding by the Flemish Government’, the Fund for
Scientific Research Flanders (FWO, project 3G004613), the
Interuniversity Attraction Poles Programme, IAP7/5,- Belgian State –
Belgian Science Policy.

References

1.     J.
Adanez, A. Abad, F. Garcia-Labiano, P. Gayan, and L.F. de Diego, Progress in
Energy and Combustion Science 38 (2012) 215

2.     L.-S.
Fan, and F. Li, Ind. Eng. Chem. Res. 49 (2010) 10200.

3.     C.
Luo, Y. Zheng, N. Ding, Q. Wu, G. Bian, and C. Zheng, Ind. Eng. Chem. Res. 49
(2010) 11778.

4.     M.
Najera, R. Solunke, T. Gardner, and G. Veser, Chemical Engineering Research and
Design 89 (2011) 1533.

5.     A.
Sim, N.W. Cant, and D.L. Trimm, Int. J. Hydrog. Energy 35 (2010) 8953-8961.

6.     S.
Abanades, Int. J. Hydrog. Energy 37 (2012) 8223.

7.     J.R.
Scheffe, M.D. Allendorf, E.N. Coker, B.W. Jacobs, A.H. McDaniel, and A.W.
Weimer, Chemistry of Materials 23 (2011) 2030.

Topics