Martin S. C.
Chan
*,
Herme G.
Baldovi
and
J. S.
Dennis
Department of Chemical Engineering and Biotechnology, University of Cambridge, CB2 3RA, UK. E-mail: mscc3@cam.ac.uk
First published on 17th January 2018
Oxygen carriers are a class of materials, typically solid oxides, that can reversibly store and release oxygen for a variety of applications in energy and chemical processes, e.g. chemical looping combustion (CLC) and chemical looping air separation (CLAS). In recent years, growing interest in these materials have been focused on their use in chemical looping selective oxidations. A method for enhancing the oxygen-carrying capacity of oxygen carriers for use in selective oxidations is presented. In this approach, one material that is selective and active in the reaction is deposited on the surface of a second material acting as a reservoir of oxygen and as a support. Here, the approach has been investigated using the selective combustion of hydrogen in the presence of ethylene, an important step in the oxidative dehydrogenation of ethane. Bismuth oxides, supported on a range of ceria–zirconia materials, were made into particulate oxygen carriers and studied for their activity, selectivity and oxygen-storage capacity. STEM-EDS imaging showed that the bismuth phase was spread uniformly over the surface of the nanocrystalline particles. XPS measurements indicated that the surface was enriched with bismuth oxide. It was found that the presence of ceria in the support supplied lattice oxygen additional to that provided by the bismuth oxide, without affecting the selectivity of bismuth oxide towards the combustion of H2. In other words, the surface chemistry was decoupled from the bulk properties of the support, thus simplifying the design and formulation of selective oxygen carriers. This demonstrates a readily-applicable generic approach for the design of oxygen carriers for selective oxidations.
Considerable research has been undertaken on the application of chemical looping for the combustion of carbonaceous fuels.9–11 The principle relies on the use of an oxygen carrier material (typically a transition metal oxide). The carrier oxidises the fuel in one reactor, becoming reduced in the process. The reduced carrier is then conveyed to a second reactor where it is reoxidised with the air reactor, before being returned to the fuel reactor to complete the cycle. This method of combustion has the key advantage of producing a pure stream of CO2, undiluted with N2, ready for capture and storage in the Earth, with minimal energy penalty compared to other separation technologies such as amine scrubbing or oxy-fuel combustion.10,12 However, the general idea of conveying solid lattice oxygen to a reaction, in the absence of gaseous oxygen, can be extended to selective oxidations. For chemical looping selective oxidation, the fuel would be substituted for some organic feed, and the flue gases would be replaced with valuable chemical products, as shown in Fig. 1.
Using chemical looping to supply oxygen has substantial advantages for selective oxidations, owing largely to the complete substitution of gaseous oxygen for lattice oxygen. In conventional oxidation using gaseous oxygen or air, the requirement to operate below the explosion limit requires substantial recycle of reactants and, should there be a lack of mixedness, presents a safety hazard, or, at least, an opportunity for partial oxidation of the feed to unwanted carbon oxides. Using chemical looping to convey the oxygen as lattice oxygen to the reaction would achieve an inherent separation between the air and organic streams, offering the advantages of (i) safety, as the risk of forming explosive mixtures is minimised, (ii) eliminating the need to operate in oxygen-lean conditions, allowing for higher single-pass conversions and decreasing the need for recycling, (iii) potentially improved selectivities, and (iv) avoiding the need to purify gaseous oxygen or the organic product from diluents such as nitrogen. In the light of these advantages, interest in chemical looping as a technology for the production of chemicals has been gradually increasing.13–18
An important example of a semi-commercial application of this technology is Dupont's process for oxidising n-butane to maleic anhydride, undertaken in a circulating fluidised bed (CFB).4 Although this process has seen significant improvement since its conception in the 1980s,19 vanadium pyrophosphate is still used as the oxygen carrier and suffers from a low oxygen-carrying capacity, of the order of 0.2 wt% (kg O/kg carrier). Oxygen carriers used for combustion typically range from 5–25 wt%. For chemical syntheses, lower capacities necessitate lower partial pressures of reactant, otherwise the carrier quickly becomes severely reduced which tends to cause coking. Lower capacities might, to some extent, be compensated for by adjusting operating parameters such as using a higher solid circulation rate (as in circulating fluidised beds) or higher frequency of cycling (as in packed bed reactors) but at an economic trade-off with higher operating costs. Higher capacities are therefore advantageous for chemical looping operation.20
In general, a suitable oxygen carrier needs to be (i) highly active, (ii) highly selective, (iii) highly stable with respect to repeated cycles of redox, (iv) have a sufficiently high oxygen-carrying capacity, and (v) be sufficiently cheap to manufacture. The problem is compounded because the selectivity and activity of the material might change substantially as its oxidation state alters with progress of the reaction (e.g. combustion may be prominent when the carrier is fully oxidised, whereas coking may occur when the carrier is deeply reduced).
For some reactions, one material may not possess all relevant properties, and so here we propose the use of structured composite materials for kinetically-controlled selective oxidations by controlling the chemical potential of oxygen supplied to the site of the reaction. The surface is decorated with selective active sites, presented by one component. This selective phase is stabilised on an active support, e.g. a solid electrolyte, that can provide additional lattice oxygen to the active sites. Lattice oxygen may be drawn from the electrolyte itself if it is reducible. The hypothesis is that gas preferentially reacts with the selective phase, which, in turn, is regenerated with oxygen from the electrolyte either by bulk diffusion or by reverse spillover at the triple phase boundary. The resulting synergistic composite is then both selective and stable, performing better than the sum of the separate components, enabled by phase cooperation.21,22 Side reactions of the gas with the electrolyte may be much slower than with the selective active sites or be physically inhibited by the selective phase acting as a barrier. This approach is easier than screening for a single material with all the required properties.
In this work, we demonstrate this proposed concept for a model reaction scheme, chosen to be the selective combustion of hydrogen in the presence of ethylene. This reaction has had some historical interest in the context of the oxidative dehydrogenation of light alkanes.5,23 This reaction was chosen because it is possible to analyse rapidly and quantitatively the gaseous products H2O and CO2. Accordingly, the reactions during the reduction step in the chemical looping cycle are reactions (1) and (2), with reoxidation described by reaction (3) (applicable for a Bi-based oxygen carrier):
![]() | (1) |
![]() | (2) |
![]() | (3) |
For the selective combustion of hydrogen in the presence of hydrocarbons, bismuth oxide24 and doped ceria1,25 have previously been demonstrated to be selective carriers. Bismuth oxide and metallic bismuth have low melting points (817 °C and 272 °C, respectively) which makes them prone to deactivation. Stabilisation with a support is therefore necessary, but this decreases the gravimetric oxygen-carrying capacity if it is inert. A possible candidate active support is ceria, which is stable and conducts oxide ions but needs to be doped to be selective. These dopants could be Bi or W.23,25 Zirconia is often mixed with ceria to increase its stability and oxygen-carrying capacity at a given temperature.26,27 A combination of bismuth oxide supported on ceria–zirconia has been shown to be active, selective and stable,6 but the effect of the support on the innate selectivity of bismuth has not been established. The effect of the composition of the support was also not established. Ceria is also interesting because it has similar lattice parameters to those of certain polymorphs of bismuth oxide, thereby promoting phase cooperation.22 The present work explores the effect of an active support on the selectivity. Accordingly, a range of ceria–zirconia supported bismuth oxides were prepared, and their performances for reactions 1 and 2 were evaluated in a packed bed reactor in chemical looping mode.
Powder X-ray diffraction (XRD) was used to analyse the crystalline phases of the oxygen carriers. Measurements were made at atmospheric conditions in a diffractometer (Empyrean, PANalytical) using Cu-Kα radiation. The reference patterns used to identify the phases were ICSD-60900 (m-ZrO2), ICSD-85322 (t-ZrO2), ICSD-88759 (CeO2), ICSD-94229 (α-Bi2O3), ICSD-417638 (β-Bi2O3), ICSD-98144 (δ-Bi2O3).
X-ray photoelectron spectrometry (XPS) was used to analyse the surface composition of the oxygen carriers. The spectrometer (ESCALAB 250Xi, Thermo Fisher Scientific) was equipped with an Al-Kα (1486.6 eV) X-ray source, and the analysis was conducted under an ultrahigh vacuum (<10−10 mbar). Binding energies were calibrated using the C(1s) peak (284.6 eV).
Nitrogen physisorption isotherms were measured at −196 °C using a Micromeritics ASAP 2020 instrument. Approximately 0.10 g of sample was used, which was degassed under vacuum at 150 °C for 1 h prior to the measurements. The isotherms were modelled using the Brunauer, Emmett and Teller (BET) method.29
Transmission electron microscopy (TEM) was used to analyse the morphology of the oxygen carriers. The samples were prepared by ultrasonically dispersing the powdered oxygen carriers in deionised water. A droplet of this transparent suspension was loaded onto holey carbon film on a 400 mesh copper grid (Agar Scientific), followed by drying in a vacuum oven at 80 °C. High-resolution TEM (HRTEM) images were obtained using a Tecnai F20 system at 200 kV with a high brightness field emission gun (FEG). High angle annular dark field (HAADF) scanning TEM (STEM) was coupled with energy-dispersive X-ray spectroscopy (EDX), performed with an Oxford EDX detector, to obtain elemental maps of the materials.
The feed gases to the reactor were supplied from cylinders (Air Liquide), and consisted of separate cylinders of (i) 5.4 vol% ethylene, (ii) 4.81 vol% H2, (iii) 5.03 vol% O2, and (iv) pure N2 (purity >99.998%). The balance gas was always N2. Gas flows were manipulated by calibrated rotameters. The gas analysis was accomplished by a set of online analysers in series. These were, in order from upstream to downstream: for H2O (IST, HYT 271, capacitive polymer humidity and temperature sensor), H2 (ABB, EL3020 Caldos27, thermal conductivity), CO, CO2 (ABB, EL3020 Uras26, NDIR), and O2 (ABB, EL3020 Magnos206, paramagnetic). A drying tube packed with CaCl2 was fitted between the humidity sensor and thermal conductivity sensor to remove moisture. Offline gas chromatography (GC) equipped with a flame ionization detector (FID) was used to measure CH4, C2H6 and C2H4 (Agilent, 6850), with separation achieved by a porous layer open tubular column (Agilent, GS-GasPro). The FID signals were calibrated using external standards from certified gas mixtures. The gas samples were withdrawn through a septum seal on the reactor exhaust line.
In the integral bed experiments, 0.10 g of the carrier, diluted with 0.80 g of SiC, was loaded into the packed bed to measure selectivity and activity. The chemical looping operation consisted of two stages: (i) reduction in a mixture of 2.5 vol% H2, 2.5 vol% C2H4, (ii) oxidation in 5 vol% O2, with nitrogen purges between stages. The operating temperature ranged from 450 to 550 °C. The reducing feed was an equimolar mixture of H2 and C2H4 to emulate the products from a catalytic dehydrogenation process. C2H6 was omitted for simplicity because it has been shown to be much less reactive than C2H4.5 Each stage progressed until no further reaction was observed. The hydrogen and carbon molar balances between the feed and the products were always within ±5%. A control was performed at 550 °C in which the reactor was packed with only SiC: no detectable conversions of H2 or C2H4 were measured. All packed bed experiments were analysed using standard diagnostic criteria for heat and mass transfer limitations: axial dispersion;30,31 bed dilution;32 particle external33 and internal mass transfer;34 particle external and internal heat transfer and bed radial heat transfer.35 It was revealed that the only limitation on the observed rates of reaction were at higher conversions due to the supply of reactant. The oxygen-carrying capacity of the materials was calculated by integrating the oxygenated products in the off-gas, using:
![]() | (4) |
For kinetic measurements, the packed bed reactor (n.b. a fluidised bed reactor was unsuitable because the prepared particles were too friable) was operated in differential mode, i.e. the conversion of the feed was less than 10% at all times, to minimise axial gradients in chemical composition. This was achieved by varying the flow rate and the mass of the carrier loaded into the bed; the residence time here was much smaller than in the integral bed experiments. Because of the short bed lengths, the oxygen carrier in these experiments was pelletised to 90–150 μm sieve diameter (with 120 grit SiC, Alfa Aesar, used as the inert diluent) to minimise the influence of bed dilution on the measured kinetics, by increasing the bed length:
particle diameter ratio.32 In a typical experiment, the bed was packed and heated to the desired temperature (in the range 450–550 °C). A cycle consisted of an oxidation step in 5 vol% O2, then a subsequent purge in N2 for 3 min, then a reduction step in 4.81 vol% H2 until no products were detected. The molar flowrate of the evolved steam during the reduction step gave a measure of the kinetics at that particular temperature. For any temperature, the cycle was repeated to check for deactivation of the carrier; deviation in the measured activity (viz. the apparent first order rate constant for the combustion of H2 at zero conversion of the carrier) was usually less than 5%. The temperature was then ramped to a higher setpoint to measure the apparent activation energy. At the end of the series of experiments, deactivation of the carrier was checked again by repeating an earlier experiment; deviation in the measured activity was usually less than 10%.
The concentration profiles of steam, from the differential bed, were deconvoluted from the response of the humidity sensor by using a first order lag model. Thus, the deconvoluted mole fraction of steam was calculated using:
![]() | (5) |
![]() | (6) |
The response time of the sensor tmix was generally greater than the characteristic reaction time by at least a factor of 5 (further details in the ESI†). Practical limitations on tmix arose because of significant pressure drops along the bed at higher flow rates – flow rates above 500 mL min−1 were not used because the pressure drop exceeded 0.2 bar.
The corrected mole fraction of steam was used to calculate the specific rate of reaction, r, at time ti, using:
![]() | (7) |
The conversion of the solid, XO, was calculated using:
![]() | (8) |
The peak conversion of H2, XH2, was calculated using:
![]() | (9) |
The peak conversion of C2H4, XC2H4, was calculated on a C-mole basis from experiments where only C2H4 was fed, using:
![]() | (10) |
The rate constant for the combustion of species i, ki, was calculated from the peak conversion Xi using:
![]() | (11) |
BET surface area (m2 g−1) | Oxygen-carrying capacity (g O/g carrier, %) | ||
---|---|---|---|
Before impregnation | After impregnation | ||
Bi25Ce0Zr | 3.7 | 2.5 | 2.44 |
Bi25Ce25Zr | 45.6 | 3.7 | 3.84 |
Bi25Ce50Zr | 39.8 | 5.2 | 3.38 |
Bi25Ce75Zr | 33.3 | 7.2 | 2.54 |
Bi25Ce100Zr | 11.1 | 7.6 | 2.66 |
Fig. 2 shows the X-ray diffraction patterns for the materials before and after impregnation with 25 wt% Bi2O3. It can be seen that the supports were apparently present in the fluorite phase (but the broad peaks meant that phases of lower symmetry cannot be excluded27), with the exception of the ZrO2 sample which presented both monoclinic and tetragonal phases. The peak positions shift with the addition of zirconia because the smaller size of the Zr4+ cation compared to the Ce4+ cation decreases the lattice parameter. Upon the addition of bismuth oxide, some samples presented additional peaks of low intensity which were attributed to α-Bi2O3. It was also found that β-Bi2O3 was present in Bi25Ce0Zr (i.e. 25 wt% Bi2O3 supported on ZrO2), consistent with previous reports.37,38 For the samples containing ceria, the low intensities of the bismuth phases could have been owing to a combination of (i) bismuth forming amorphous layers not detectable by diffraction, (ii) the doping of bismuth into the ceria or ceria–zirconia lattice, or (iii) broad overlapping peaks. There is also evidence of shifted peak positions upon the addition of Bi2O3, most apparent for the pure CeO2 support, suggesting doping of the fluorite structure.
Table 2 summarises the near-surface (depth ≲ 10 nm) elemental compositions of the oxygen carriers, as measured by XPS. There was a deviation in the XPS measurements from the mean composition, because XPS is a surface-sensitive technique. It can be seen that (a) the surface is enriched with Bi, as expected from the impregnation procedure, and also (b) the Ce:
Zr ratio is less than the bulk ratio, in agreement with other workers.6Table 2 also shows the results for Bi25Ce75Zr reduced in H2, showing a lower amount of Bi than in the freshly-calcined sample; this has also been reported by other workers.6
XPS measurements | Mean composition | |||||
---|---|---|---|---|---|---|
Bi mol% | Ce mol% | Zr mol% | Bi mol% | Ce mol% | Zr mol% | |
Bi25Ce0Zr | 32 | 0 | 68 | 15.0 | 0 | 85.0 |
Bi25Ce25Zr | 35 | 5.2 | 59 | 16.2 | 20.9 | 62.8 |
Bi25Ce50Zr | 38 | 11 | 51 | 17.4 | 41.3 | 41.3 |
Bi25Ce75Zr | 47 | 23 | 31 | 18.6 | 61.0 | 20.3 |
Bi25Ce100Zr | 69 | 31 | 0 | 19.8 | 80.2 | 0 |
Reduced Bi25Ce75Zr | 19 | 33 | 48 | 18.6 | 61.0 | 20.3 |
TEM images were obtained for Bi25Ce50Zr. In Fig. 3a, it can be seen that the samples were composed of nanocrystalline particles. Fringe patterns were also visible, with an interplanar spacing of 0.30 nm corresponding to the (111) plane of ceria–zirconia. Under these ex situ conditions, amorphous material can also be seen on the surface. Fig. 3b shows EDS maps, confirming that Bi was homogeneously dispersed over the surface of each particle. The mean metal composition in Fig. 3b was found to be: 40% Ce, 37% Zr and 24% Bi. This differed from Table 2 because EDS is not as surface-sensitive because the incident electrons penetrate to a much greater depth, whereas photoelectrons in XPS can only escape from near the surface. This means that the XPS measurement is more relevant in this context of heterogeneous catalysis.
The oxygen carriers differed in their activities and in their oxygen-carrying capacities. Fig. 5 shows that as the Ce content of the support increased, the conversion of H2 increased. Also shown are the conversions of C2H4 from a separate experiment where only 5.4 vol% C2H4 was fed to the reactor. The conversions were significantly higher than in the case where H2 was co-fed with the ethylene. This shows that the activity towards the combustion of C2H4 was strongly dependent on the composition of the gas. Nonetheless, the degree of oxidation of the ethylene is substantially less than that of H2, as seen in Fig. 5. Fig. 6 shows that the oxygen-carrying capacity of each sample increased with temperature, and is higher for the intermediate ratios of Ce:
Zr, consistent with existing studies.26 Bi25Ce0Zr displayed a capacity consistent with its loading of 25 wt% Bi2O3, i.e. 0.25 × 10.3 = 2.58 wt%, which again confirms the loading of bismuth (pure Bi2O3 has a capacity of 10.3 wt%). The supports were able to contribute additional lattice oxygen with the bismuth oxide without adversely affecting the selectivity, despite ceria–zirconia itself not being selective.6,25 The capacity of a bismuth-free sample, Ce0.75Zr0.25O2, was found to be 1.40 wt% at 550 °C, so the expected oxygen-capacity of the composite carrier with the addition of 25 wt% Bi2O3, assuming an ideal mixture, would then be (0.75 × 1.40) + (0.25 × 10.3) = 3.63 wt%. This is consistent with the measured value shown in Fig. 6 of 3.67 wt%, which confirms that the oxygen-carrying capacity of the ceria–zirconia phase is not affected by the addition of bismuth. Furthermore, the selectivity of the oxygen contained in the support was enhanced by the presence of a selective surface; this was apparent when comparing the amount of CO2 evolved per mass of support with and without Bi2O3, which were 18 and 34 [μmol CO2]/[g Ce0.75Zr0.25O2], respectively, despite both supports releasing the same total amount of oxygen. Lastly, the sample with pure ceria as the support did not contribute much additional lattice oxygen, despite being the most active; the lack of zirconia meant that bulk ceria was less reducible.
![]() | ||
Fig. 7 Variation of the specific rate of reaction with the partial pressure of hydrogen, shown for Bi25Ce75Zr. |
Fig. 8 shows how the rates of reaction varied with the conversion of the oxygen carriers. In all samples except Bi25Ce100Zr, an initial period of constant rate of reaction can be observed for up to 30% conversion of the solid. As the solid conversion increases, the rate falls because the reaction becomes starved of lattice oxygen. The overall form of the curves of r against XO is largely similar. The response of the sensor became more significant for the most active samples, Bi25Ce75Zr and Bi25Ce100Zr, indicated by the maxima being reached at higher values of XO (i.e. greater than ∼10%). The possible influence of the response of the sensor on the measured kinetics is examined further in the ESI.†
The maximum rates of reaction in Fig. 8 were used to calculate the apparent first order rate constants according to eqn (11). These rate constants are shown in the Arrhenius plot in Fig. 9 and in Table 3. These linear fits only use T ≤ 500 °C for Bi25Ce75Zr and T ≤ 475 °C for Bi25Ce100Zr because of convolution with the sensor at the higher temperatures for these samples. It can be seen that the apparent activation energies do not vary significantly for any of the samples (with the exception of Bi25Ce75Zr), which suggests a similar rate-determining step. The similar selectivities in these bismuth-containing samples, also shown in Table 3, also suggest a common active site.
Enhanced oxygen-carrying capacities were observed for the samples with a mixture of ceria and zirconia as the support. This enhancement, in combination with consistently high values of selectivity, demonstrates a synergistic effect between the selective bismuth oxide phase and the ceria–zirconia phase. The cyclic stability and oxygen-carrying capacity of ceria–zirconia complemented the selectivity and capacity of bismuth oxide, resulting in a composite material that performed better than either material in isolation. This synergy may be described as ‘phase cooperation’22 and is likely to have been facilitated by similar lattice parameters. The lattice parameters for the potentially relevant structures are similar; 5.4112 Å (fluorite), 5.6446 Å (tetragonal, c-axis), and 5.6549 Å (fluorite), for CeO2, β-Bi2O3, and δ-Bi2O3, respectively. In general, this could be readily applied in the design of selective oxygen carriers for chemical syntheses. This presents a solution to previous attempts at a chemical looping arrangement, which have, in some cases, been beset by poor oxygen-carrying capacities.20,39
Footnote |
† Electronic supplementary information (ESI) available: Calculations of rates of heat and mass transfer, and their qualification with standard diagnostic criteria for the measurement of kinetics. Limitation of the sensor response time on the measured kinetics. Calculation of oxygen-free composition from XPS. Data supporting this work are available from http://www.repository.cam.ac.uk. See DOI: 10.1039/c7cy01992k |
This journal is © The Royal Society of Chemistry 2018 |