Skip to main content
Erschienen in: Journal of Materials Science 23/2020

Open Access 27.03.2020 | Advanced Nano Materials

Co-precipitation synthesized nanostructured Ce0.9Ln0.05Ag0.05O2−δ materials for solar thermochemical conversion of CO2 into fuels

verfasst von: Gorakshnath Takalkar, Rahul R. Bhosale, Fares AlMomani, Suliman Rashid

Erschienen in: Journal of Materials Science | Ausgabe 23/2020

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

Synthesis, characterization, and application of Ce0.9Ln0.05Ag0.05O2−δ materials (where, Ln = La, Pr, Nd, Sm, Gd, Tb, Dy, Er) for the thermochemical conversion of CO2 reported in this paper. The Ce0.9Ln0.05Ag0.05O2−δ materials were synthesized by using an ammonium hydroxide-driven co-precipitation method. The derived Ce0.9Ln0.05Ag0.05O2−δ materials were characterized via powder X-ray diffraction, scanning electron microscope, and electron diffraction spectroscopy. The characterization results indicate the formation of spherically shaped Ce0.9Ln0.05Ag0.05O2−δ nanostructured particles. As-prepared Ce0.9Ln0.05Ag0.05O2−δ materials were further tested toward multiple CO2 splitting cycles by utilizing a thermogravimetric analyzer. The results obtained indicate that all the Ce0.9Ln0.05Ag0.05O2−δ materials produced higher quantities of O2 and CO than the previously studied pure CeO2 and lanthanide-doped ceria materials. Overall, the Ce0.911La0.053Ag0.047O1.925 showed the maximum redox reactivity in terms of O2 release (72.2 μmol/g cycle) and CO production (136.6 μmol/g cycle).
Hinweise

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
\( n_{{{\text{O}}_{2} }} \)
Moles of O2 (μmol/g)
\( n_{\text{CO}} \)
Moles of CO (μmol/g)
\( \Delta m_{\text{loss}} \)
Amount of loss in the mass (mg)
\( \Delta m_{\text{gain}} \)
Amount of gain in the mass (mg)
\( M_{{{\text{O}}_{2} }} \)
Molecular weight of O2 (g/mol)
\( M_{\text{O}} \)
Molecular weight of O (g/mol)
\( m_{{{\text{Ce}}_{0.9} {\text{Ln}}_{0.05} {\text{Ag}}_{0.05} {\text{O}}_{ 2 { - \delta }} }} \)
Mass of the \( {\text{Ce}}_{0.9} {\text{Ln}}_{0.05} {\text{Ag}}_{0.05} {\text{O}}_{{2{{ - \delta }}}} \) materials (mg)

Introduction

The worldwide human population is rising at a quick pace, which inherently demands a large amount of energy for consumption [1]. Current energy production largely depends on the utilization of fossil fuels. The excessive use of petroleum-based resources results in a continuous CO2 discharge [2, 3]. This constant release of CO2 is considered as one of the primary reasons for the variation in the climate parameters [4]. The environmental distress occurring due to increases in CO2 emission generated more interest in the conversion of CO2 into value-added products.
One of the possible options for CO2 utilization is to convert the captured CO2 into fuels. Solar thermochemical cycles driven based on the metal oxide (MO)-based redox reactions can split CO2 into CO (Fig. 1) [5, 6]. The solar CO produced can be combined with the solar H2 (produced via solar thermochemical splitting of water) for the manufacturing of the solar syngas, which can be further utilized in the catalytic Fischer–Tropsch process [7].
The MO-based solar-driven thermochemical conversion of CO2 is a two-phase process. In the first phase, the MO is reduced with the help of concentrated solar power, and in phase 2, the reduced MO is re-oxidized again via a CO2 splitting reaction. Zinc oxide [8, 9], tin oxide [10, 11], iron oxide [12, 13], CeO2 [1416], doped ceria [1720], ferrites [2124], perovskites [2528], and others [2932] have been utilized for the solar thermochemical conversion of H2O and CO2. Among all these, the phase pure CeO2 is considered as one of the promising options due to its faster reaction kinetics and better thermal stability. Although the CeO2 is beneficial for the thermochemical conversion of H2O and CO2, a lower fuel production capacity is one of the major limitations associated with this MO.
Recently, Bhosale and Takalkar [33] reported that the doping of lanthanides such as La, Pr, Nd, Sm, Gd, Tb, Dy, and Er into CeO2 fluorite cubic crystal structure improved the thermal reduction (TR) capacity of the CeO2. It was also reported that although the TR capability of CeO2 was improved, only Ce0.9La0.1O2 was capable of producing higher quantities of CO than CeO2 via CO2 splitting reaction. In another investigation, Takalkar et al. [19] reported that the inclusion of Ag in the transition metal-doped ceria considerably improved the TR as well as CO2 splitting (CDS) capacity.
Based on the results reported in our previous investigations, in this study, the synthesis, characterization, and application of Ce0.9Ln0.05Ag0.05O2−δ materials (where, Ln = La, Pr, Nd, Sm, Gd, Tb, Dy, Er) for the thermochemical conversion of CO2 is reported. Synthesis of the Ce0.9Ln0.05Ag0.05O2−δ materials is achieved via a co-precipitation method. The derived materials are further tested for multiple thermochemical CDS cycles by utilizing a thermogravimetric analyzer (TGA) setup. The TR and CDS capacity of the Ce0.9Ln0.05Ag0.05O2−δ materials was estimated and compared with the previously studied phase pure CeO2 and lanthanide-doped ceria materials.

Experimental

Preparation and characterization of Ce0.9Ln0.05Ag0.05O2−δ

Nitrate-based precursors of ceria, silver, and all lanthanides were acquired from Sigma-Aldrich, USA. Aqueous 28% NH3OH as a precipitating agent was procured from the same supplier. Ultrapure deionized (DI) water (produced through Direct-Q system, Millipore, France) was utilized for the preparation of nitrate solution. An ultrapure grade Ar gas (purity 99.999%) and 50% CO2 + 50% Ar gas mixture are ordered from the Buzwair Scientific and Technical Gases, Doha, Qatar.
The synthesis of redox Ce0.9Ln0.05Ag0.05O2−δ materials was achieved via co-precipitation of the hydroxide method. As shown in Fig. 2, an aqueous mixture of selected metal precursors was prepared by dissolving them in 300 ml of deionized water. To this mixture, aqueous ammonium hydroxide (28%) was added to attain a pH approximately equal to 10. The mixture further stirred for 24 h with a maintained pH ~ 10. Once the stirring was stopped, the precipitate of Ce0.9Ln0.05Ag0.05O2−δ material was allowed to settle via gravity (mixture kept undisturbed overnight). The next morning, the supernatant liquid was decanted, and the obtained precipitate was washed several times by deionized water with the help of a vacuum filtration unit. The obtained filtered cake of Ce0.9Ln0.05Ag0.05O2−δ was dried (120 °C, 5 h), crushed, and annealed (Nabertherm Furnace) up to 1000 °C for 4 h in the presence of air. The obtained annealed powder was analyzed for the determination of the phase/elemental composition and morphology by using Panalytical XPert powder X-ray diffractometer (PXRD) and scanning electron microscope (SEM, Nova Nano 450, FEI) equipped with the electron diffraction spectroscopy (EDS).

CO2 splitting experiments

The Ce0.9Ln0.05Ag0.05O2−δ materials were experimentally tested in a TGA setup (Setaram Instrumentation, France), which is shown in Fig. 3. The experimental parameters utilized to perform the thermochemical cycles are given in Table 1. Approximately, 50 mg of the calcined Ce0.9Ln0.05Ag0.05O2−δ powder was charged in an alumina (100 µl) crucible, and then placed inside the heating furnace of the TGA. Before performing the TGA experiments, the residual air filling the hollow space near to the furnace was evacuated by applying a vacuum followed by sweeping by the inert Ar. Chilled water (generated by Julabo FC 1600T) was utilized to decrease the exiting gas stream temperature. Additional details related to the TGA setup and the experimental procedure are already described in our previous studies [23, 29]. Multiple TR and CDS steps were performed by considering the operating parameters given in Table 1.
Table 1
Experimental conditions used for the TGA experiments
Parameters
Value
Carrier gas
Ultrapure inert Ar
Feed gas mixture
50% CO2 + 50% Ar
Flowrate of carrier and feed gas mixture
100 ml/min
TR and CDS temperature
1400 °C and 1000 °C
Heating and cooling rate
25 K/min
Reaction time for TR and CDS
60 min and 30 min
The mass variations allied with the Ce0.9Ln0.05Ag0.05O2−δ materials during the TR and CDS steps were documented after subtracting the blank TGA experimental data from the actual TGA experimental data. The amount of O2 liberated (µmol/g) during the TR step and the quantity of CO produced during the CDS step are calculated by utilizing Eqs. (1) and (2).
$$ n_{{{\text{O}}_{2} }} = \frac{{\Delta m_{\text{loss}} }}{{\left( {M_{{{\text{O}}_{2} }} \times m_{{{\text{Ce}}_{0.9} {\text{Ln}}_{0.05} {\text{Ag}}_{0.05} {\text{O}}_{2{ - \delta }} }}} \right)}} $$
(1)
$$ n_{\text{CO}} = \frac{{\Delta m_{gain} }}{{\left( {M_{\text{O}} \times m_{{{\text{Ce}}_{ 0. 9} {\text{Ln}}_{ 0. 0 5} {\text{Ag}}_{ 0. 0 5} {\text{O}}_{{ 2 {{ - \delta }}}} }} } \right)}} $$
(2)

Results and discussion

After synthesizing the Ce0.9Ln0.05Ag0.05O2−δ materials, the next important step was to determine the phase composition of the derived materials. PXRD profiles of the calcined Ce0.9Ln0.05Ag0.05O2−δ materials are shown in Fig. 4a, b. The presented patterns indicate a cubic fluorite crystal structure of the Ce0.9Ln0.05Ag0.05O2−δ materials, similar to the one reported in the case of CeO2 [23]. The PXRD profiles shown in Fig. 4a further indicates the absence of the formation of any metal or metal oxide impurities. As shown in Fig. 4b, the Ce0.9Ln0.05Ag0.05O2−δ material peaks shift toward either lower or higher 2θ angle (based on the crystal ionic radii of the dopants). This shift in the peaks further confirmed the successful incorporation of the dopants inside the ceria fluorite crystal structure. The formation of nominally phase pure Ce0.9Ln0.05Ag0.05O2−δ materials was also assured via EDS analysis (results given in Table 2).
Table 2
Abbreviations, chemical composition, crystallite size, and cell parameter of each Ce0.9Ln0.05Ag0.05O2−δ material
Abbreviations
Ce0.9Ln0.05Ag0.05O2−δ composition (via EDS)
Crystallite size (nm)
Cell parameter (Å)
La5Ag5Ce
Ce0.911La0.053Ag0.047O1.925
63.1
5.434
Pr5Ag5Ce
Ce0.892Pr0.051Ag0.051O1.886
46.6
5.411
Nd5Ag5Ce
Ce0.921Nd0.049Ag0.050O1.940
32.3
5.417
Sm5Ag5Ce
Ce0.896Sm0.048Ag0.053O1.890
60.2
5.421
Gd5Ag5Ce
Ce0.889Gd0.054Ag0.050O1.884
39.4
5.389
Tb5Ag5Ce
Ce0.905Tb0.051Ag0.046O1.909
41.3
5.410
Dy5Ag5Ce
Ce0.910Dy0.053Ag0.048O1.923
61.5
5.411
Er5Ag5Ce
Ce0.899Er0.050Ag0.055O1.900
38.1
5.411
In order to determine the crystallite size, a widely used Scherrer equation and the PXRD data associated with the Ce0.9Ln0.05Ag0.05O2−δ materials were utilized. Table 2 indicates that the variation in the crystallite size of the Ce0.9Ln0.05Ag0.05O2−δ materials does not follow any specific trend. As per the obtained results, the Ce0.9Ln0.05Ag0.05O2−δ materials can be arranged as follows based on their average crystallite size: La5Ag5Ce > Dy5Ag5Ce > Sm5Ag5Ce > Pr5Ag5Ce > Tb5Ag5Ce > Gd5Ag5Ce > Er5Ag5Ce > Nd5Ag5Ce.
After estimating the composition and crystallite size of each Ce0.9Ln0.05Ag0.05O2−δ material, the morphology was scrutinized by performing the SEM analysis. The SEM images obtained looks very similar to each other and indicate the formation of roughly spherical particles of Ce0.9Ln0.05Ag0.05O2−δ. The microscopic SEM analysis further confirmed that the particles were agglomerated. It was also observed that the average particle size was very close to the crystallite sizes given in Table 2. The representative SEM images of Pr5Ag5Ce, Tb5Ag5Ce, Gd5Ag5Ce, and Er5Ag5Ce are shown in Fig. 5.
The redox performance of the Ce0.9Ln0.05Ag0.05O2−δ materials is estimated by performing the thermochemical CDS experiments using the TGA setup. Ce0.9Ln0.05Ag0.05O2−δ materials thermally reduced at 1400 °C (10 K/min) for 60 min in the presence of the inert Ar (100 ml/min) and then re-oxidized at 1000 °C by using a feed gas mixture containing 50% CO2 + 50% Ar (100 ml/min). As an example, Fig. 6 represents a TGA profile of La5Ag5Ce material obtained during the first CDS cycle. As shown in Fig. 6, during the TR step, the mass of the La5Ag5Ce material reduced by 0.511 mg, and during the CDS step, the weight of the La5Ag5Ce material increased by 0.126 mg. These mass variations further converted into respective redox performances in terms of \( n_{{{\text{O}}_{2} }} \) released (320.2 µmol/g) and \( n_{\text{CO}} \) produced (157.8 µmol/g) by using Eqs. (1) and (2).
The mass variations associated with the Ce0.9Ln0.05Ag0.05O2−δ materials recorded during the first cycle are shown in Fig. 7a, b. The \( n_{{{\text{O}}_{2} }} \) released and \( n_{\text{CO}} \) produced by each Ce0.9Ln0.05Ag0.05O2−δ material was computed based on the obtained TGA profiles and given in Table 3. The data given in Table 3 show that the Pr5Ag5Ce was capable of releasing a higher amount of O2 at 1400 °C than the other Ce0.9Ln0.05Ag0.05O2−δ materials. Likewise, the CO production aptitude of La5Ag5Ce was the uppermost when compared with the remaining Ce0.9Ln0.05Ag0.05O2−δ materials. The numbers listed in the \( n_{\text{CO}} /n_{{{\text{O}}_{2} }} \) ratio column shows that the re-oxidation ability of the Nd5Ag5Ce was better than the other Ce0.9Ln0.05Ag0.05O2−δ materials.
Table 3
Redox performance of Ce0.9Ln0.05Ag0.05O2−δ materials during the first cycle
Ce0.9Ln0.05Ag0.05O2−δ
\( n_{{{\text{O}}_{2} }} \) released (µmol/g)
\( n_{\text{CO}} \) produced (µmol/g)
\( n_{\text{CO}} /n_{{{\text{O}}_{2} }} \) ratio
La5Ag5Ce
320.0
157.8
0.49
Pr5Ag5Ce
548.6
152.7
0.28
Nd5Ag5Ce
191.0
153.5
0.80
Sm5Ag5Ce
195.3
125.9
0.64
Gd5Ag5Ce
326.5
125.0
0.38
Tb5Ag5Ce
212.3
139.7
0.66
Dy5Ag5Ce
365.8
143.9
0.39
Er5Ag5Ce
415.4
134.1
0.32
Interesting to note that the \( n_{\text{CO}} \) produced by each Ce0.9Ln0.05Ag0.05O2−δ material was considerably lower than the \( n_{{{\text{O}}_{2} }} \) released during the first cycle. The two probable reasons for these results are (a) poor re-oxidation ability of the Ce0.9Ln0.05Ag0.05O2−δ materials or (b) additional mass loss during the first TR reduction due to the release of volatile chemicals from the Ce0.9Ln0.05Ag0.05O2−δ materials (which remained unburnt during the calcination step). For further investigation of this matter, the Ce0.9Ln0.05Ag0.05O2−δ materials were tested for three cycles (by maintaining the same operating conditions utilized in the case of the first cycle). Figure 8 shows the variations in the mass of the Ce0.9Ln0.05Ag0.05O2−δ materials during the successive three thermochemical cycles. Besides, Fig. 9a and b compares the \( n_{{{\text{O}}_{2} }} \) released and \( n_{\text{CO}} \) produced by each Ce0.9Ln0.05Ag0.05O2−δ material from cycle 1 to cycle 3.
Figures 8 and 9 show that the \( n_{{{\text{O}}_{2} }} \) released by all the Ce0.9Ln0.05Ag0.05O2−δ materials during cycle 1 was considerably higher than cycle 2. For example, the \( n_{{{\text{O}}_{2} }} \) released by La5Ag5Ce, Nd5Ag5Ce, Gd5Ag5Ce, and Dy5Ag5Ce in cycle 2 was lower by 77.9%, 64.6%, 76.6%, and 80.0% as compared to cycle 1. The comparison between cycle 2 and cycle 3 shows a different story. The \( n_{{{\text{O}}_{2} }} \) released by all the Ce0.9Ln0.05Ag0.05O2−δ materials in cycle 3 was slightly less when compared to cycle 2. For instance, the \( n_{{{\text{O}}_{2} }} \) released by La5Ag5Ce, Nd5Ag5Ce, Gd5Ag5Ce, and Dy5Ag5Ce in cycle 3 was lower by 2.0%, 0.0%, 14.8%, and 1.8% than cycle 2. Based on the results given in Figs. 8 and 9, it can be concluded that the prime reason for the more substantial O2 evolution in cycle 1 was the additional loss in the mass of the Ce0.9Ln0.05Ag0.05O2−δ materials due to the release of volatile chemicals.
In the case of the CDS step, the \( n_{\text{CO}} \) produced by each Ce0.9Ln0.05Ag0.05O2−δ material first decreased in cycle 2 (compared to cycle 1) and remained approximately stable in cycle 3 (compared to cycle 2). For example, the \( n_{\text{CO}} \) produced by La5Ag5Ce, Nd5Ag5Ce, Gd5Ag5Ce, and Dy5Ag5Ce in cycle 2 was 11.1%, 20.0%, 7.4%, and 24.5% lower than cycle 1. In contrast, the %decrease in the \( n_{\text{CO}} \) produced by La5Ag5Ce, Nd5Ag5Ce, Gd5Ag5Ce, and Dy5Ag5Ce in cycle 3 was dropped to 1.1%, 0.5%, 0.9%, and 1.2% when compared with the cycle 2 data. The \( n_{\text{CO}} /n_{{{\text{O}}_{2} }} \) ratio of all the Ce0.9Ln0.05Ag0.05O2−δ materials increased significantly in cycle 2 when compared with cycle 1. For instance, the \( n_{\text{CO}} /n_{{{\text{O}}_{2} }} \) ratio of La5Ag5Ce, Nd5Ag5Ce, Gd5Ag5Ce, and Dy5Ag5Ce increased by 1.49, 1.01, 1.13, and 1.09 in cycle 2 when compared with cycle 1. The \( n_{\text{CO}} /n_{{{\text{O}}_{2} }} \) ratio of all the Ce0.9Ln0.05Ag0.05O2−δ materials in cycle 2 and cycle 3 were almost identical.
The results obtained during cycle 3 indicate that most of the Ce0.9Ln0.05Ag0.05O2−δ materials are moving toward achieving a stable redox reactivity. For attaining further confirmation about the stable redox reactivity, the Ce0.9Ln0.05Ag0.05O2−δ materials were tested for ten successive cycles. The TGA profiles associated with the ten cycles are shown in Fig. 10. It was already confirmed that the data obtained in cycle 1 is misleading, and hence the TGA analysis was more focused on the remaining nine cycles. The \( n_{{{\text{O}}_{2} }} \) released and \( n_{\text{CO}} \) produced by each Ce0.9Ln0.05Ag0.05O2−δ material from cycle 2 to cycle 10 are shown in Figs. 11 and 12.
As per the data given in Fig. 11, the La5Ag5Ce, Pr5Ag5Ce, and Nd5Ag5Ce shows a stable release of O2 from cycle 2 to cycle 10. The Gd5Ag5Ce indicates redox stability in terms of constant O2 release from cycle 3 to cycle 10. For the rest of the Ce0.9Ln0.05Ag0.05O2−δ materials, i.e., Sm5Ag5Ce, Tb5Ag5Ce, Dy5Ag5Ce, and Er5Ag5Ce, a steady \( n_{{{\text{O}}_{2} }} \) evolution was realized after cycle 5 or cycle 6. In terms of average \( n_{{{\text{O}}_{2} }} \) released from cycle 2 to cycle 10, La5Ag5Ce (72.2 μmol of O2/g cycle) and Tb5Ag5Ce (60.2 μmol of O2/g cycle) were observed to be the best and worst among all the Ce0.9Ln0.05Ag0.05O2−δ materials.
As shown in Fig. 12, the La5Ag5Ce, Pr5Ag5Ce, Nd5Ag5Ce, Gd5Ag5Ce, and Dy5Ag5Ce showed a stable production of CO from cycle 2 to cycle 10. In contrast, a constant \( n_{\text{CO}} \) production in the case of Sm5Ag5Ce, Tb5Ag5Ce, and Er5Ag5Ce was noticed from cycle 6 to cycle 10. In terms of the average \( n_{\text{CO}} \) produced from cycle 2 to cycle 10, the Ce0.9Ln0.05Ag0.05O2−δ materials can be arranged as: La5Ag5Ce > Nd5Ag5Ce ~ Pr5Ag5Ce > Gd5Ag5Ce > Tb5Ag5Ce > Sm5Ag5Ce ~ Er5Ag5Ce > Dy5Ag5Ce. According to the data given in Fig. 13, the re-oxidation ability (average \( n_{\text{CO}} \) / \( n_{{{\text{O}}_{2} }} \) ratio) of the La5Ag5Ce and Tb5Ag5Ce was the highest as compared to the rest of the Ce0.9Ln0.05Ag0.05O2−δ materials. Based on \( n_{{{\text{O}}_{2} }} \) released and \( n_{\text{CO}} \) produced from cycle 2 to cycle 10, the La5Ag5Ce appears to be the most excellent candidate among all the Ce0.9Ln0.05Ag0.05O2−δ materials investigated in this study.
Table 4 reports the comparison of Ce0.9Ln0.05Ag0.05O2−δ materials with the CeO2 and Ce0.9Ln0.1O2 materials. The results given in Table 4 shows that all the Ce0.9Ln0.05Ag0.05O2−δ materials were capable of higher \( n_{{{\text{O}}_{2} }} \) release (except for Sm5Ag5Ce) and \( n_{\text{CO}} \) production than CeO2 and their corresponding Ce0.9Ln0.1O2 materials. For instance, the \( n_{{{\text{O}}_{2} }} \) released by La5Ag5Ce, Pr5Ag5Ce, Nd5Ag5Ce, Gd5Ag5Ce, Tb5Ag5Ce, Dy5Ag5Ce, and Er5Ag5Ce was higher by 21.8 μmol of O2/g cycle, 23.1 μmol of O2/g cycle, 23.0 μmol of O2/g cycle, 18.6 μmol of O2/g cycle, 0.20 μmol of O2/g cycle, 12.1 μmol of O2/g cycle, and 8.50 μmol of O2/g cycle as compared to Ce0.9La0.1O2, Ce0.9Pr0.1O2, Ce0.9Nd0.1O2, Ce0.9Gd0.1O2, Ce0.9Tb0.1O2, Ce0.9Dy0.1O2, and Ce0.9Er0.1O2, respectively. Similarly, the La5Ag5Ce, Pr5Ag5Ce, Nd5Ag5Ce, Sm5Ag5Ce, Gd5Ag5Ce, Tb5Ag5Ce, Dy5Ag5Ce, and Er5Ag5Ce produced 1.39, 1.44, 1.52, 1.32, 1.49, 1.43, 1.29, and 1.37 times higher CO when compared with the Ce0.9La0.1O2, Ce0.9Pr0.1O2, Ce0.9Nd0.1O2, Ce0.9Sm0.1O2, Ce0.9Gd0.1O2, Ce0.9Tb0.1O2, Ce0.9Dy0.1O2, and Ce0.9Er0.1O2, respectively. The overall results of this investigation indicate that the incorporation of Ag in the Ln-doped ceria was beneficial to improve the redox performance of Ce0.9Ln0.05Ag0.05O2−δ materials.
Table 4
Comparison between the pure CeO2−δ, Ce0.9Ln0.1O2−δ, and Ce0.9Ln0.05Ag0.05O2−δ [19,33]
Materials
\( n_{{{\text{O}}_{2} }} \) released (μmol/g cycle)
\( n_{\text{CO}} \) produced (μmol/g cycle)
CeO2
47.7
94.9
La5Ag5Ce
72.2
136.6
Ce0.9La0.1O2
50.4
98.0
Pr5Ag5Ce
69.7
124.3
Ce0.9Pr0.1O2
46.6
86.2
Nd5Ag5Ce
66.5
124.5
Ce0.9Nd0.1O2
43.5
82.1
Sm5Ag5Ce
61.8
113.3
Ce0.9Sm0.1O2
69.9
85.8
Gd5Ag5Ce
65.3
115.4
Ce0.9Gd0.1O2
46.7
77.2
Tb5Ag5Ce
60.2
114.0
Ce0.9Tb0.1O2
60.0
80.0
Dy5Ag5Ce
65.3
112.6
Ce0.9Dy0.1O2
53.2
87.3
Er5Ag5Ce
70.5
113.3
Ce0.9Er0.1O2
62.0
82.5

Summary and conclusions

The PXRD and EDS analysis have confirmed the formation of nominally phase pure Ce0.9Ln0.05Ag0.05O2−δ materials via co-precipitation of the hydroxide method. The average crystallite size of the derived Ce0.9Ln0.05Ag0.05O2−δ materials was in the range of 32–64 nm. The SEM analysis further established a spherical nanostructured particle morphology of the Ce0.9Ln0.05Ag0.05O2−δ materials. The SEM analysis also indicates that the doping of the lanthanides and silver does not have any significant effect on the morphology of Ce0.9Ln0.05Ag0.05O2−δ materials. Based on the results associated with the TGA analysis, the redox reactivity of the Ce0.9Ln0.05Ag0.05O2−δ materials can be ranked as: Ce0.911La0.053Ag0.047O1.925 > Ce0.921Nd0.049Ag0.050O1.940 ~ Ce0.892Pr0.051Ag0.051O1.886 > Ce0.889Gd0.054Ag0.050O1.884 > Ce0.905Tb0.051Ag0.046O1.909 > Ce0.896Sm0.048Ag0.053O1.890 ~ Ce0.899Er0.050Ag0.055O1.900 > Ce0.910Dy0.053Ag0.048O1.923. The TGA results also indicate that the Ce0.911La0.053Ag0.047O1.925 and Ce0.905Tb0.051Ag0.046O1.909 possess the highest re-oxidation ability as compared to the rest of the Ce0.9Ln0.05Ag0.05O2−δ materials. The overall results of this investigation indicate that the inclusion of Ag into the Ln-doped ceria considerably improved the O2 releasing and CO production capability of the Ce0.9Ln0.05Ag0.05O2−δ materials as compared to the phase pure CeO2 and Ce0.9Ln0.1O2 materials. After estimating the CO production capacity, our research team is currently focused on the production of H2 by using the Ce0.9Ln0.05Ag0.05O2−δ materials.

Acknowledgments

Open Access funding provided by the Qatar National Library. This publication was made possible by the NPRP Grant (NPRP8-370-2-154) from the Qatar National Research Fund (a member of Qatar Foundation). The statements made herein are solely the responsibility of author(s). The authors also gratefully acknowledge the Center of Advances Materials (CAM) at Qatar University for carrying out XRD analysis and the Central Laboratory Unit (CLU) for services related to scanning electron microscopy.

Compliance with ethical standards

Conflict of interest

The authors declare that they have no conflict of interest.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
1.
Zurück zum Zitat Varghese OK, Paulose M, LaTempa TJ, Grimes CA (2009) High-rate solar photocatalytic conversion of CO2 and water vapor to hydrocarbon fuels. Nano Lett 9:731–737CrossRef Varghese OK, Paulose M, LaTempa TJ, Grimes CA (2009) High-rate solar photocatalytic conversion of CO2 and water vapor to hydrocarbon fuels. Nano Lett 9:731–737CrossRef
2.
Zurück zum Zitat Patil GN, Vaidya PD, Kenig EY (2011) Reaction kinetics of CO2 in aqueous methyl-and dimethylmonoethanolamine solutions. Ind Eng Chem Res 51:1592–1600CrossRef Patil GN, Vaidya PD, Kenig EY (2011) Reaction kinetics of CO2 in aqueous methyl-and dimethylmonoethanolamine solutions. Ind Eng Chem Res 51:1592–1600CrossRef
3.
Zurück zum Zitat Bhosale RR, Mahajani VV (2013) Kinetics of thermal degradation of renewably prepared amines useful for flue gas treatment. J Renew Sustain Energy 5:063110CrossRef Bhosale RR, Mahajani VV (2013) Kinetics of thermal degradation of renewably prepared amines useful for flue gas treatment. J Renew Sustain Energy 5:063110CrossRef
4.
Zurück zum Zitat Anderson TR, Hawkins E, Jones PD (2016) CO2, the greenhouse effect and global warming: from the pioneering work of Arrhenius and Callendar to today’s Earth System Models. Endeavour 40:178–187CrossRef Anderson TR, Hawkins E, Jones PD (2016) CO2, the greenhouse effect and global warming: from the pioneering work of Arrhenius and Callendar to today’s Earth System Models. Endeavour 40:178–187CrossRef
5.
Zurück zum Zitat Carrillo RJ, Scheffe JR (2017) Advances and trends in redox materials for solar thermochemical fuel production. Sol Energy 156:3–20CrossRef Carrillo RJ, Scheffe JR (2017) Advances and trends in redox materials for solar thermochemical fuel production. Sol Energy 156:3–20CrossRef
6.
Zurück zum Zitat Koepf E, Alxneit I, Wieckert C, Meier A (2017) A review of high temperature solar driven reactor technology: 25 years of experience in research and development at the Paul Scherrer Institute. Appl Energy 188:620–651CrossRef Koepf E, Alxneit I, Wieckert C, Meier A (2017) A review of high temperature solar driven reactor technology: 25 years of experience in research and development at the Paul Scherrer Institute. Appl Energy 188:620–651CrossRef
7.
Zurück zum Zitat Dry ME (2002) The fischer–tropsch process: 1950–2000. Catal Today 71:227–241CrossRef Dry ME (2002) The fischer–tropsch process: 1950–2000. Catal Today 71:227–241CrossRef
8.
Zurück zum Zitat Bhosale RR (2018) Thermodynamic efficiency analysis of zinc oxide based solar driven thermochemical H2O splitting cycle: effect of partial pressure of O2, thermal reduction and H2O splitting temperatures. Int J Hydrog Energy 43:14915–14924CrossRef Bhosale RR (2018) Thermodynamic efficiency analysis of zinc oxide based solar driven thermochemical H2O splitting cycle: effect of partial pressure of O2, thermal reduction and H2O splitting temperatures. Int J Hydrog Energy 43:14915–14924CrossRef
9.
Zurück zum Zitat Abanades S, Charvin P, Flamant G (2007) Design and simulation of a solar chemical reactor for the thermal reduction of metal oxides: case study of zinc oxide dissociation. Chem Eng Sci 62:6323–6333CrossRef Abanades S, Charvin P, Flamant G (2007) Design and simulation of a solar chemical reactor for the thermal reduction of metal oxides: case study of zinc oxide dissociation. Chem Eng Sci 62:6323–6333CrossRef
10.
Zurück zum Zitat Bhosale RR, Kumar A, Sutar P (2017) Thermodynamic analysis of solar driven SnO2/SnO based thermochemical water splitting cycle. Energy Convers Manag 135:226–235CrossRef Bhosale RR, Kumar A, Sutar P (2017) Thermodynamic analysis of solar driven SnO2/SnO based thermochemical water splitting cycle. Energy Convers Manag 135:226–235CrossRef
11.
Zurück zum Zitat Chambon M, Abanades S, Flamant G (2011) Thermal dissociation of compressed ZnO and SnO2 powders in a moving-front solar thermochemical reactor. AIChE J 57:2264–2273CrossRef Chambon M, Abanades S, Flamant G (2011) Thermal dissociation of compressed ZnO and SnO2 powders in a moving-front solar thermochemical reactor. AIChE J 57:2264–2273CrossRef
12.
Zurück zum Zitat Scheffe JR et al (2011) Hydrogen production via chemical looping redox cycles using atomic layer deposition-synthesized iron oxide and cobalt ferrites. Chem Mater 23:2030–2038CrossRef Scheffe JR et al (2011) Hydrogen production via chemical looping redox cycles using atomic layer deposition-synthesized iron oxide and cobalt ferrites. Chem Mater 23:2030–2038CrossRef
13.
Zurück zum Zitat Cho WC et al (2012) Reactivity of iron oxide as an oxygen carrier for chemical-looping hydrogen production. Int J Hydrog Energy 37:16852–16863CrossRef Cho WC et al (2012) Reactivity of iron oxide as an oxygen carrier for chemical-looping hydrogen production. Int J Hydrog Energy 37:16852–16863CrossRef
14.
Zurück zum Zitat Chueh WC, Haile SM (2010) A thermochemical study of ceria: exploiting an old material for new modes of energy conversion and CO2 mitigation. Philos Trans A Math Phys Eng Sci 368:3269–3294CrossRef Chueh WC, Haile SM (2010) A thermochemical study of ceria: exploiting an old material for new modes of energy conversion and CO2 mitigation. Philos Trans A Math Phys Eng Sci 368:3269–3294CrossRef
15.
Zurück zum Zitat Bhosale RR et al (2019) A decade of ceria based solar thermochemical H2O/CO2 splitting cycle. Int J Hydrog Energy 44:34–60CrossRef Bhosale RR et al (2019) A decade of ceria based solar thermochemical H2O/CO2 splitting cycle. Int J Hydrog Energy 44:34–60CrossRef
16.
Zurück zum Zitat Ackermann S, Scheffe JR, Steinfeld A (2014) Diffusion of oxygen in ceria at elevated temperatures and its application to H2O/CO2 splitting thermochemical redox cycles. J Phys Chem C 118:5216–5225CrossRef Ackermann S, Scheffe JR, Steinfeld A (2014) Diffusion of oxygen in ceria at elevated temperatures and its application to H2O/CO2 splitting thermochemical redox cycles. J Phys Chem C 118:5216–5225CrossRef
17.
Zurück zum Zitat Scheffe JR, Jacot R, Patzke GR, Steinfeld A (2013) Synthesis, characterization, and thermochemical redox performance of Hf4, Zr4, and Sc3 doped ceria for splitting CO2. J Phys Chem C 117:24104–24114CrossRef Scheffe JR, Jacot R, Patzke GR, Steinfeld A (2013) Synthesis, characterization, and thermochemical redox performance of Hf4, Zr4, and Sc3 doped ceria for splitting CO2. J Phys Chem C 117:24104–24114CrossRef
18.
Zurück zum Zitat Le Gal A, Abanades S (2011) Catalytic investigation of ceria-zirconia solid solutions for solar hydrogen production. Int J Hydrog Energy 36:4739–4748CrossRef Le Gal A, Abanades S (2011) Catalytic investigation of ceria-zirconia solid solutions for solar hydrogen production. Int J Hydrog Energy 36:4739–4748CrossRef
19.
Zurück zum Zitat Takalkar G et al (2018) Transition metal doped ceria for solar thermochemical fuel production. Sol Energy 172:204–211CrossRef Takalkar G et al (2018) Transition metal doped ceria for solar thermochemical fuel production. Sol Energy 172:204–211CrossRef
20.
Zurück zum Zitat Muhich CL et al (2013) Efficient generation of H2 by splitting water with an isothermal redox cycle. Science 341:540–542CrossRef Muhich CL et al (2013) Efficient generation of H2 by splitting water with an isothermal redox cycle. Science 341:540–542CrossRef
21.
Zurück zum Zitat Shende RV, Puszynski JA, Opoku MK, Bhosale RR (2009) Synthesis of novel ferrite foam material for water-splitting application. In: Proceedings of NSTIdNanotech conference & expo., ISBN, 2009 Shende RV, Puszynski JA, Opoku MK, Bhosale RR (2009) Synthesis of novel ferrite foam material for water-splitting application. In: Proceedings of NSTIdNanotech conference & expo., ISBN, 2009
22.
Zurück zum Zitat Bhosale RR, Shende RV, Puszynski JA (2012) Sol-gel derived NiFe2O4 modified with ZrO2 for hydrogen generation from solar thermochemical water-splitting reaction. In: MRS online proceedings library archive, vol 1387 Bhosale RR, Shende RV, Puszynski JA (2012) Sol-gel derived NiFe2O4 modified with ZrO2 for hydrogen generation from solar thermochemical water-splitting reaction. In: MRS online proceedings library archive, vol 1387
23.
Zurück zum Zitat Agrafiotis C, Zygogianni A, Pagkoura C, Kostoglou M, Konstandopoulos AG (2013) Hydrogen production via solar-aided water splitting thermochemical cycles with nickel ferrite: experiments and modeling. AIChE J 59:1213–1225CrossRef Agrafiotis C, Zygogianni A, Pagkoura C, Kostoglou M, Konstandopoulos AG (2013) Hydrogen production via solar-aided water splitting thermochemical cycles with nickel ferrite: experiments and modeling. AIChE J 59:1213–1225CrossRef
24.
Zurück zum Zitat Scheffe JR, Li J, Weimer AW (2010) A spinel ferrite/hercynite water-splitting redox cycle. Int J Hydrog Energy 35:3333–3340CrossRef Scheffe JR, Li J, Weimer AW (2010) A spinel ferrite/hercynite water-splitting redox cycle. Int J Hydrog Energy 35:3333–3340CrossRef
25.
Zurück zum Zitat Scheffe JR, Weibel D, Steinfeld A (2013) Lanthanum–strontium–manganese perovskites as redox materials for solar thermochemical splitting of H2O and CO2. Energy Fuels 27:4250–4257CrossRef Scheffe JR, Weibel D, Steinfeld A (2013) Lanthanum–strontium–manganese perovskites as redox materials for solar thermochemical splitting of H2O and CO2. Energy Fuels 27:4250–4257CrossRef
26.
Zurück zum Zitat Dey S, Naidu B, Govindaraj A, Rao C (2015) Noteworthy performance of La1–x Cax MnO3 perovskites in generating H2 and CO by the thermochemical splitting of H2 O and CO2. Phys Chem Chem Phys 17:122–125CrossRef Dey S, Naidu B, Govindaraj A, Rao C (2015) Noteworthy performance of La1–x Cax MnO3 perovskites in generating H2 and CO by the thermochemical splitting of H2 O and CO2. Phys Chem Chem Phys 17:122–125CrossRef
27.
Zurück zum Zitat Bhosale RR et al (2017) La-based perovskites as oxygen-exchange redox materials for solar syngas production. MRS Adv 2:3365–3370CrossRef Bhosale RR et al (2017) La-based perovskites as oxygen-exchange redox materials for solar syngas production. MRS Adv 2:3365–3370CrossRef
28.
Zurück zum Zitat Gálvez M et al (2015) Physico-chemical changes in Ca, Sr and Al-doped La–Mn–O perovskites upon thermochemical splitting of CO2 via redox cycling. Phys Chem Chem Phys 17:6629–6634CrossRef Gálvez M et al (2015) Physico-chemical changes in Ca, Sr and Al-doped La–Mn–O perovskites upon thermochemical splitting of CO2 via redox cycling. Phys Chem Chem Phys 17:6629–6634CrossRef
29.
Zurück zum Zitat Bhosale RR, Kumar A, AlMomani F, Ghosh U, Khraisheh M (2017) A comparative thermodynamic analysis of samarium and erbium oxide based solar thermochemical water splitting cycles. Int J Hydrog Energy 42:23416–23426CrossRef Bhosale RR, Kumar A, AlMomani F, Ghosh U, Khraisheh M (2017) A comparative thermodynamic analysis of samarium and erbium oxide based solar thermochemical water splitting cycles. Int J Hydrog Energy 42:23416–23426CrossRef
30.
Zurück zum Zitat Bhosale R et al (2016) Solar hydrogen production via a samarium oxide-based thermochemical water splitting cycle. Energies 9:316CrossRef Bhosale R et al (2016) Solar hydrogen production via a samarium oxide-based thermochemical water splitting cycle. Energies 9:316CrossRef
31.
Zurück zum Zitat Bhosale RR et al (2016) Solar hydrogen production via erbium oxide based thermochemical water splitting cycle. J Renew Sustain Energy 8:034702CrossRef Bhosale RR et al (2016) Solar hydrogen production via erbium oxide based thermochemical water splitting cycle. J Renew Sustain Energy 8:034702CrossRef
33.
Zurück zum Zitat Bhosale R, Takalkar G (2018) Nanostructured co-precipitated Ce0.9Ln0.1O2 (Ln = La, Pr, Sm, Nd, Gd, Tb, Dy, or Er) for thermochemical conversion of CO2. Ceram Int 44:16688–16697CrossRef Bhosale R, Takalkar G (2018) Nanostructured co-precipitated Ce0.9Ln0.1O2 (Ln = La, Pr, Sm, Nd, Gd, Tb, Dy, or Er) for thermochemical conversion of CO2. Ceram Int 44:16688–16697CrossRef
Metadaten
Titel
Co-precipitation synthesized nanostructured Ce0.9Ln0.05Ag0.05O2−δ materials for solar thermochemical conversion of CO2 into fuels
verfasst von
Gorakshnath Takalkar
Rahul R. Bhosale
Fares AlMomani
Suliman Rashid
Publikationsdatum
27.03.2020
Verlag
Springer US
Erschienen in
Journal of Materials Science / Ausgabe 23/2020
Print ISSN: 0022-2461
Elektronische ISSN: 1573-4803
DOI
https://doi.org/10.1007/s10853-020-04567-w

Weitere Artikel der Ausgabe 23/2020

Journal of Materials Science 23/2020 Zur Ausgabe

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.