Next Article in Journal
Solvent-Free Mechanochemical Preparation of Metal-Organic Framework ZIF-67 Impregnated by Pt Nanoparticles for Water Purification
Next Article in Special Issue
Bio-Adipic Acid Production from Muconic Acid Hydrogenation on Palladium-Transition Metal (Ni and Zn) Bimetallic Catalysts
Previous Article in Journal
New Magnetically Assembled Electrode Consisting of Magnetic Activated Carbon Particles and Ti/Sb-SnO2 for a More Flexible and Cost-Effective Electrochemical Oxidation Wastewater Treatment
Previous Article in Special Issue
Ni5P4-NiP2-Ni2P Nanocomposites Tangled with N-Doped Carbon for Enhanced Electrochemical Hydrogen Evolution in Acidic and Alkaline Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influences of Co-Content on the Physico-Chemical and Catalytic Properties of Perovskite GdCoxFe1−xO3 in CO Hydrogenation

by
Elizaveta M. Borodina
1,
Liliya V. Yafarova
2,
Tatiana A. Kryuchkova
1,
Tatiana F. Sheshko
1,*,
Alexander G. Cherednichenko
1 and
Irina A. Zvereva
2,*
1
Department of Physical and Colloidal Chemistry, Faculty of Science, Peoples’ Friendship University of Russia (RUDN University), 6 Miklukho-Maklaya Street, 117198 Moscow, Russia
2
Department of Chemical Thermodynamics and Kinetics, Institute of Chemistry, Saint Petersburg State University, 7/9 Universitetskaya nab., 199034 Saint Petersburg, Russia
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(1), 8; https://doi.org/10.3390/catal13010008
Submission received: 27 November 2022 / Revised: 16 December 2022 / Accepted: 19 December 2022 / Published: 22 December 2022
(This article belongs to the Special Issue Transition Metal Complexes as Catalysts)

Abstract

:
The effect of the substitution of cobalt into the GdFeO3 perovskite structure on the selective hydrogenation of CO was investigated. A series of GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) samples were synthesized by sol-gel technology and characterized by XRD, BET specific area, DSC, TG, EDX and XPS. The experimental data made it possible to reveal a correlation between the state of iron and cobalt atoms, the fractions of surface and lattice oxygen, and catalytic characteristics. It has been found that varying the composition of GdCoxFe1−xO3 complex oxides leads to a change in the oxygen-metal binding energy in Gd-O-Me, the ratio of metals in various oxidation states, and the amount of surface and lattice oxygen, which affects the adsorption and catalytic characteristics of complex oxides.

Graphical Abstract

1. Introduction

Light olefins are among the basic petrochemical products. Currently, the global production of these products is estimated at 185 million tons [1], and the global production of propylene already exceeds 100 million tons per year [2]. The continuous growth in demand for light olefins is largely determined by the increase in the consumption of polyethylene and polypropylene and the expansion of their technological applications. To solve this problem, it is necessary to create a variety of effective catalytic technologies to produce light olefins from available raw materials.
One of the promising methods of ethylene production is the process of hydrogenation of CO. This process implies the use of noble metal-based catalysts such as Ru, Pt, Pd and Rh because of their inherent resistance to carbon deposition on the surface. However, the use of these catalysts is not economically viable. The industry also prefers nickel and cobalt catalysts because these metals are inexpensive and readily available. However, nickel- and cobalt-based catalysts are more prone to coke deposition under these synthesis conditions, resulting in reduced activity and deactivation [3,4,5].
Thus, the development of low-budget, effective and sintering-resistant catalysts for obtaining ethylene is an important problem. The issue of studying and developing new catalytic systems, improving already used processes and creating new methods aimed at obtaining olefins under conditions of hydrogenation of carbon monoxide remains relevant nowadays. Oxides with a perovskite structure (ABO3) are considered an alternative to traditional catalysts for many industrial processes due to their high thermal stability, the ability to control physicochemical properties and low cost [6,7,8]. One of the advantages of perovskite-like oxides is the possibility of creating a wide range of compositions by complete/partial, as well as isovalent/non-isovalent substitution of A and B cations while maintaining the structure. This allows for controlling the stability of the structure [9], electronic [10], redox and surface properties of complex oxides [11], creating new functional materials [12,13,14,15,16,17,18,19,20].
Perovskites have good flexibility and diversity in their chemical composition and can accommodate solid defects, such as vacancies at both the cation and anion sites [21], [22]. The B-site transition metals on the surface are believed to be the active centers owing to the exposed 3d electron orbitals. Considering that the B-site transition metals embedded in the perovskite lattice are atomically dispersed, we can expect to develop highly active and anti-coking catalysts.
Perovskites have been widely studied in oxidation reactions [23,24,25]. However, the study of Fe- and Co-containing perovskites in the hydrogenation of carbon monoxide is insufficiently described in the literature [26,27,28,29,30,31]. From the analysis of the existing works, it can be concluded that there is a close relationship between the nature and chemical state of the elements responsible for the activity and the type of surface C-containing intermediates and product distributions. Accordingly, A(Fe/Co)O3 perovskite-type oxides with reducible B3+ ions under typical reaction conditions are excellent model catalysts for gaining new and useful insight into the surface process taking place during CO activation and carbon chain growth.
Thereby the purpose of this work was to study the catalytic activity of systems based on perovskite-type iron- and cobalt-containing complex oxides in the process of carbon monoxide hydrogenation and to establish the relationship between the physicochemical properties of complex oxides and their catalytic characteristics. It should be noted that catalysts of similar composition worked well for dry reforming of methane [32] and diesel soot oxidation [33].

2. Results and Discussion

2.1. Characterization

Complex oxides GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) were studied by X-ray phase analysis, X-ray photoelectron spectroscopy and SEM+EDX. Detailed XFA and XPS results are presented in [26]. All samples are single-phase and have an orthorhombic structure with the spatial group Pbnm and Pnma. All diffraction peaks can be attributed to the orthorhombic structure of GdFeO3 (PDF-ICDD 01-072-9908), GdCoO3 (PDF-ICDD 00-025-1057) and their solid solutions.
Analysis of the dependence of the structural parameters a, b and c on the composition of GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) (Figure 1) shows that the values of parameters a and c decrease slightly with a greater iron substitution for cobalt in ferrite, while the lattice parameter b remains practically unchanged for all complex oxides except for the ferrite samples with cobalt fraction x = 0.8 and completely substituted GdCoO3. This indicates that the lattice geometry distorts with increasing cobalt content.
The crystallite size of the obtained samples estimated using the Scherrer formula is in the range from ~47 to 65 nm (Table 1) and does not show a dependence on the composition [33].
The specific surface area of the obtained compounds was determined by the BET method (Table 1); for GdFeO3, the highest value was observed, which is consistent with the results of scanning electron microscopy. For gadolinium cobaltite and solid solutions GdCoxFe1−xO3 (x = 0.2; 0.5; 0.8), the low values of the specific surface area are due to the relatively high calcination temperature (800 °C). At the same time, the values of the surface area of the obtained oxides were found to be higher than for similar compounds prepared by the ceramic method (less than 1.0 m2/g) [34]. Thus, the values of the specific surface area do not show a dependence on the composition, and the conditions of preparation have a great influence.
The surface morphology of gadolinium ferrite and cobaltite, as well as their solid solutions obtained after calcining at 800 °C for 1 h, was studied using scanning electron microscopy. Figure 2 shows photomicrographs of the surface of oxides with the composition GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1); all studied compounds demonstrate a similar surface morphology, which indicates that the complete or partial substitution of iron for cobalt in the perovskite structure does not affect surface morphology.
The micrographs of the surface in Figure 2 show a uniform morphology without particle agglomeration. The surface is characterized by a “porous” structure, possibly associated with combustion and gas evolution during oxide production. It should be noted that, in contrast to gadolinium cobaltite and partially substituted ferrites, gadolinium ferrite (GdFeO3) exhibits a more porous morphology, which is due to a lower preparation temperature and is consistent with the results of determining the specific surface area (Table 1).
To study the composition of the surface, X-ray spectral analysis was used and elemental mapping of the surface was carried out. The results show that the amount of Gd and Fe/Co on the surface of the obtained compounds is close to the theoretical value (Table 1).
XPS was used to study the surface composition, the oxidation state and the distribution of atoms in the obtained compounds. Traditional photoelectron spectroscopy is a surface-sensitive method in which the depth of probing in the energy range 200 eV–1500 eV does not exceed 2 nm. Figure 3 presents the overview XPS spectra of oxides GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) in the binding energy range of 0–1400 eV and represents the intense peaks of Gd3d, Fe2p, Co2p, O1s and C1s (signals of other elements were not found). The chemical states of the atoms in the compounds under study were refined by decomposing the obtained spectra described in [33].
The partial substitution of iron for cobalt in the samples GdCoxFe1−xO3 (x = 0.2; 0.5; 0.8) resulted in the heterovalent state Fe2+, Fe3+(711.0; 724.0 eV) and Co2+/Co3 (779.97; 794.0 eV) (Table 2). The heterovalent state causes the appearance of oxygen vacancies in the structure of Co-containing samples and changes in the bond lengths Me-O-Me [35].
The O1s photoelectron spectra of compounds GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) shown in Figure 4 are asymmetric and wide, which confirms the presence of various types of oxygen in these compounds [36,37]. Analysis and decomposition of the spectra showed that the obtained compounds contained various types of oxygen (at binding energies of ~533.00 eV, 530.90 and 528.57–529.67 eV), which are referred to as physically absorbed oxygen (H2O), chemically adsorbed oxygen (Os:O2, O22− and O) and lattice oxygen (Ol), respectively. Table 2 presents a quantitative assessment of various types of oxygen in the studied oxides. The Os/Ol ratio in GdCoxFe1−xO3 at x = 0.2; 0.5; 0.8; 1 increases with an increase in the amount of cobalt, and these four samples have almost the same values of the surface area. At the same time, for the GdFeO3 (x = 0), a higher value of the surface area and a higher value of the Os/Ol ratio are observed. The study [38] shows that chemisorbed oxygen (Os) plays an important role in low-temperature oxidative reactions, while lattice oxygen (Ol) affects the oxidizing ability of compounds in high-temperature reactions. Thus, the Os/Ol ratio can be used to assess the oxidative characteristics of the compounds under study.
As was shown in [33], when using the H2-TPR and O2-TPD methods, the mobility of lattice oxygen increases with increasing cobalt content in GdCoxFe1−xO3 samples. This tendency confirmed that there are a large number of cationic vacancies in the Co-rich samples due to the easier Co3+→Co2+ transition.

2.2. CO Hydrogenation

The catalytic activity in the reaction of carbon monoxide hydrogenation was evaluated in the temperature range 523–708 K at a CO:H2 ratio = 1:2 (CO conversions are shown in Figure 5). The substitution of small amounts of cobalt (x < 0.5) led to the suppression of CO conversions compared to the initial gadolinium ferrite. On the sample GdCo0.2Fe0.8O3 (x = 0.8), degrees of conversion of carbon monoxide did not exceed 45%. Substitution of iron for cobalt (x = 0.2; 0.5) leads to an improvement in oxygen mobility and a decrease in the amount of surface oxygen Os. This is reflected in the decrease in CO conversion as CO is adsorbed through the surface oxygen, but the introduction of more cobalt into ferrite gadolinium (x > 0.5) promoted the growth of CO conversion and it reached 90–95%.
For GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8) samples, the conversions barely depended on temperature and reached almost maximum values at 523 K. This is due to the participation in the reaction of carbon particles formed because of the dissociative adsorption of CO at lower temperatures [21]. A slight increase in the CO conversion with increasing temperature may be due to the involvement of not only surface oxygen but lattice oxygen as well. The amount of CO in the gas phase correlated with its conversions and was determined by the type of catalyst. At the same time, on all catalysts, the process was accompanied by the formation of carbon dioxide, the amounts of which also depended on the composition of perovskite (Figure 6).
The amount of formed carbon dioxide in the region up to 623 K was approximately the same, but further increasing the temperature led to an increase in the rate of CO2 formation. It should be noted that on the sample GdCoO3, the process of CO2 formation was much more intense than on the other samples.
It is known that CO adsorption mainly occurs at the A-centers of perovskite with the formation of carbonate complexes Gd2O2CO3 [39]. The decomposition of these complexes produces carbon dioxide (Equation (1)).
Gd2O2CO3(s) → CO2(g) + Gd2O3(s)
In the studied perovskites, the parameters a and c of the crystal lattice change; hence, the oxygen-metal bond energy in Gd-O-Me will also change, which in turn leads to a change in the gadolinium-oxygen bond and is reflected in the value of CO adsorption. The formation of CO2 is also possible in the interaction of the adsorbed COads molecule with perovskite (OS) oxygen (Equation (2)):
COads + OS → CO2
As stated above, the substitution of iron for cobalt improves oxygen mobility and redox properties and changes the ratio of surface and bulk oxygen in the structure of complex oxides: the share of Os decreases, while the share of Ol increases. It can be assumed that with increasing temperatures, both surface and lattice oxygen begin to participate in the reaction (2) (the amount of lattice oxygen depends on the composition of the oxide, i.e., with cobalt enrichment the oxidation capacity of the samples increases).
The products of hydrogenation of carbon monoxide on GdCoxFe1−xO3 catalysts (x = 0; 0.2; 0.5; 0.8; 1) were C1–C5 hydrocarbons, and the main ones were methane, ethylene, ethane and propylene. The rate of formation of hydrocarbons began at 573 K and increased with increasing temperatures.
A comparison of the rates of formation of the main products (methane, ethylene, propylene and CO2) showed that the substitution of Fe in gadolinium ferrite for Co leads to an increase in the specific catalytic activity (Figure 7), but on the GdCo0.2Fe0.8O3, a decrease in the rate of product formation was observed. It is worth noting that as the proportion of cobalt in the initial ferrite increases, the formation of propylene stops. The formation rates of methane and ethylene increased in the series:
GdCo0.2Fe0.8O3 ˂ GdFeO3 ˂ GdCo0.5Fe0.5O3 ˂ GdCoO3 ˂ GdCo0.8Fe0.2O3
The appearance of a synergistic effect on GdCo0.8Fe0.2O3 is due to the fact that this sample has different structural characteristics from the others.
Varying the composition of catalysts led to a change in the quantitative ratio of products. Thus, when performing the reaction on GdFeO3, the content of methane, ethylene and propylene at T = 648 K was 76%, 14% and 6%, respectively. The complete substitution of iron for cobalt increased the amount of methane to 85%. However, the GdCo0.2Fe0.8O3 sample showed a decrease in methane to 38%, ethylene to 9% and propylene to 5%. The sample with the cobalt substitution fraction x = 0.8 had an ethylene content of 13%.
The dependence of the hydrocarbon chain growth factor αi on the number of carbon atoms (i) in the chain of intermediate products is shown in Figure 8.
Regardless of the catalyst composition, the highest values were for α 1 , i.e., the formation of hydrocarbons with two carbon atoms was most likely in the studied systems, and on the GdCo0.2Fe0.8O3 catalyst, this probability reached a maximum of 50%.
From the literature data [35] it is known that iron hydrogen is adsorbed mainly in atomic form and molecular forms are practically absent, but cobalt hydrogen can be adsorbed in both atomic and molecular forms. Moreover, on the iron centers, hydrogen is more mobile relative to the surface, unlike cobalt, because in the series Fe, Co and Ni, bond energy Me-O grows [25,35]. Thus, increasing the proportion of cobalt in the B-position leads to the growth of the molecular forms of hydrogen and their ordering. The probability of the formation of CH2 particles increases and results in the formation of hydrocarbons with an even number of carbon atoms.
The calculation of selectivity for target products (ethylene, propylene and butylene) showed that the substitution of iron for cobalt in the perovskite led to a decrease in selectivity for the target reaction products. Thus, at 573 K, the C 2 = - C 4 = selectivity on GdFeO3 was 29% and 20% on GdCoO3 (Figure 9). The replacement of iron with a small amount of cobalt led to an increase (of approximately 32%) in the selectivity of light olefins. The most selective with respect to ethylene turned out to be the sample with a fraction of cobalt x = 0.2, for which the Co2+/∑ Con+ ratio is maximal and Fe2+/∑ Fen+ is minimal.
To determine kinetic characteristics, such as the effective activation energies of product formation and the logarithms of the pre-exponential multiplier (an indirect characteristic of the number of the active centers of the catalyst surface), the interpolation of experimental data in linear coordinates of the Arrhenius equation was performed. On the Arrhenius dependences of the rates of formation of ethylene for samples with a content of cobalt x > 0.5 two areas of linearity with a transition temperature of about T* ≈ 598–609.7 K were observed. The presence of two linear sections with different values of the activation energies Ea indicates either a change in the state of the active Men+ centers depending on the temperature (change in the oxidation state of Co3+ ↔ Co2+) or the course of different processes of a complex reaction in each temperature interval (T < T* is chemisorption of reactants, and T > T* is the conversion of chemisorbed complexes into products). The calculated values of the effective activation energies for the formation of reaction products and the logarithms of the prefactor for all the catalysts studied are shown in Table 3.
As the data in Table 3 show, the introduction of a small amount of cobalt led to a decrease in the activation energies of ethylene formation, while the value of the pre-exponential multiplier also decreased. The activation energies of methane formation and the logarithms of the prefactor were comparable to those of the original ferrite. For the GdCo0.8Fe0.2O3 sample, the effective activation energies were the highest. However, in spite of the high activation energy, high values of ln K0 are also observed. For the same sample, the minimum Co2+ content can be seen. For samples with a fraction of cobalt x ≥ 0.5, there are comparable values of Os/Ol and Fe2+/(Fe2+ + Fe3+), but only the samples with x = 0.8 the minimum Co2+/∑ Con+ ratio and the highest rates of product formation were observed. At the same time, the most selective was the sample with the cobalt fraction x = 0.2, for which the Co2+/∑ Con+ ratio is maximal and Fe2+/∑ Fen+ is minimal.
It can be assumed that Co3+ and Fe3+ in the Gd-O-Me bond are active centers, and the presence of Co2+ somewhat complicates the process and shifts it towards the formation of olefins.
The determination of the deposited carbon on the surface of the catalytic systems because of the reaction was carried out by thermogravimetric analysis (TG) and differential scanning calorimetry (DSC). Figure 10 shows the mass loss curves for GdCoxFe1−xO3 samples (x = 0; 0.5; 1) as an example. In the temperature range of 553–593 K, a slight increase in the mass of the GdCo0.5Fe0.5O3 sample was observed, which is most likely due to the process of oxygen chemisorption from the air environment on the surface carbon and the subsequent decrease in mass with exo-effects in the 573–753 K range is due to carbon desorption in the form of CO2.
TG analysis showed a loss of mass of the samples up to temperatures of ~893–913 K (620–640 °C). The amount of deposited carbon increases with increasing cobalt content. The largest amount was observed in GdCoO3. However, for all investigated perovskite systems, the formation of the so-called “resin” as well as the “encapsulation” of the catalyst particles was not observed.

3. Materials and Methods

3.1. Catalyst Preparation

The synthesis of GdCoxFe1−xO3 complex oxides (x = 0; 0.2; 0.5; 0.8; 1) was performed using the citrate-nitrate sol-gel technique [40,41]. Stoichiometric Gd(NO3)3·6H2O (99.9%, Vecton, St. Petersburg, Russia), Fe(NO3)3·9H2O (98%, Vecton, St. Petersburg, Russia) и Co(NO3)3·9H2O (98.5%, Vecton, St. Petersburg, Russia) nitrates were dissolved in distilled water, and then the flask was placed on a magnetic stirrer with a thermostat. Citric acid was added to the resulting salt solution with constant stirring. After the complete dissolution of citric acid, to set the pH of the solution at 6, an ammonia solution was added, and then the solution was evaporated at a temperature of ~120 °C until the resulting gel ignited and a black powder was formed. Then the temperature was raised to 450 °C and the resulting powder was calcined for 2 h. The samples containing cobalt were additionally calcined at 800 °C for 1 h.

3.2. Characterization

Powder X-ray diffraction (XRD) was used to determine the phase composition and structural data of the obtained compounds. The analysis was carried out using a Rigaku MiniFlex II diffractometer with CuKα radiation. Diffraction patterns of all investigated compounds were obtained under the same conditions sufficient to detect all characteristic reflections: the range of angles is 2θ = 10–60°, and the scanning speed is 5°/min. The obtained diffraction patterns were analyzed by comparing the present reflections with the data of the ICDD-PDF2 database. The data obtained allowed us to conclude the phase composition of the studied compounds.
X-ray photoelectron spectroscopy (XPS) was used to analyze the surface composition and determine the state of atoms. The measurements were carried out on a Thermo Fisher Scientific Escalab 250Xi spectrometer using radiation AlKα = 1486.6 eV, spectral resolution-0.5 eV. The characteristic C1s line (C-C bond), which is observed at a bond energy of 284.8 eV, was used as a control line. Experiments on temperature-programmed hydrogen reduction (H2-TPR) were carried out on a Micrometrics AutoChem II 2920 analyzer. Samples with a mass of <0.05 g were placed in a U-shaped quartz reactor and heated with a constant flow of a gas mixture (10% H2/90% Ar, feed rate = 50 mL/min) up to 1000 °C (heating rate = 10 °C/min). The amount of absorbed gas was measured using a thermal conductivity detector (TCD). In order to exclude the effect of water vapor on the operation of the TCD, a trap was placed in front of the detector and in a Dewar with ethyl alcohol (T = −139 °C). Sample temperature, valve position and gas flow rate data were monitored and controlled using a computer equipped with Micromeritics AutoChem II software.
Thermogravimetric analysis using a Netzsch STA 449 F5 Jupiter was performed to estimate the amount of deposited carbon after catalytic tests. The studies were carried out on samples with a suspension of 20–30 mg in a stream of air (rate = 50 mL/min−1) at a heating rate of 10 K/min−1 in the temperature range 303–1173 K. The specific surface area was determined by the method of low-temperature nitrogen adsorption at T = 77 K on Nova 4200e (Quantachrome, Boynton Beach, FL, USA) and QuadrasorbSI devices. The samples were preliminarily degassed at 300 °C for 5 h. The obtained adsorption-desorption isotherms were used to estimate the specific surface area of the samples by applying the BET method.

3.3. Catalytic Activity Tests

Catalyst performance of the samples in the carbon monoxide hydrogenation reaction was evaluated in a quartz tubular microreactor containing 0.1 g of the sample diluted in 0.5 g of quartz in order to avoid the formation of hot spots in the catalytic bed alongside increasing the volume. Activity tests were carried out in the temperature range 523–708 K under atmospheric pressure with the feed composition CO:H2 = 1:2 and total flow rate of 1.5 L⋅h−1, which corresponds to GHSV of 8700 h−1. Gaseous effluent was analyzed online using a gas chromatograph instrument (Crystal 5000.2, a column of stainless steel filled with Porapack Q, argon as a carrier gas) fitted with thermal conductivity and flame ionization detectors.
The catalytic characteristics were calculated using the following equations:
X i ,   % = n i n t n o u t n i n t × 100
S i = R i R i · 100 % ,
R i = n i o u t ω V m
where nint и nout is the initial amount of substance and mole yield of products, ∑ni is the mole yield of hydrocarbon reaction products (mole), Ri is the rate of formation of the i-th reaction product per 1 g catalyst (mol/h⋅g), ∑Ri is the sum of yields of the i-th reaction products per 1 g catalyst (mol/h⋅g), ω is volumetric rate of the reaction mixture (L/h), V is the volume of the chromatograph loop (0.153 × 10−3 L) and m is the mass of the catalyst, g.
The carbon chain growth factor αi was calculated using Equation (7) [42]:
α i = k > i Y k k k i Y k k
where α i is carbon chain growth factor for the intermediate with the number of carbon atoms I, Yk is the yield of the component with the number of carbon atoms k.
The coefficient α1 is the probability of the addition of CO to the intermediate containing one carbon atom to form an intermediate with two carbon atoms; α2 is the probability of the next step of the addition of CO to the intermediate with two carbon atoms, etc.

4. Conclusions

Complex perovskite-type oxides GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1.0) prepared by the sol-gel method were studied as catalysts for the hydrogenation of carbon monoxide. The effect of iron substitution by cobalt on the physicochemical and catalytic properties in the process was investigated. It was found that the metals in the samples were in heterovalent states Fe2+/Fe3+ and Co2+/Co3+, which is compensated by oxygen vacancies. It was found that the partial substitution of iron for cobalt in GdCoxFe1−xO3 (0.5 < x < 1) and a decrease in the Co2+ share in the sample resulted in an increase in CO conversions; nevertheless, x = 0.2 with the maximum Co2+/∑ Con+ ratio and minimum Fe2+/∑ Fen+ ratio was the most selective for ethylene. It has been suggested that Co3+ and Fe3+ are active centers, while the presence of Co2+ somewhat complicates the process and shifts it towards the formation of olefins. It was also found that varying the composition of GdCoxFe1−xO3 complex oxides leads to changes in the oxygen-metal bond energy in Gd-O-Me, the ratio of metals in different oxidation states, and the amount of surface and lattice oxygen, which is reflected in the adsorption and catalytic characteristics of complex oxides and change of the quantitative ratio of olefins in the process of CO hydrogenation.

Author Contributions

Conceptualization and methodology, I.A.Z. and T.F.S.; investigation, E.M.B., T.A.K. and L.V.Y.; data curation, T.A.K.; writing—original draft, E.M.B., L.V.Y. and T.A.K. writing—review and editing, T.A.K., T.F.S. and I.A.Z.; supervision, A.G.C.; project administration, A.G.C. All authors have read and agreed to the published version of the manuscript.

Funding

This publication was supported by the RUDN University Scientific Projects Grant System, project № 021521-2-174 and the scholarship of the President of the Russian Federation (№ CП-686.2021.1).

Data Availability Statement

The original data are available from E.M.B. and L.V.Y.

Acknowledgments

Authors are also grateful to Saint Petersburg State University Research Park. The XRD study was carried out at the Research Center for X-ray Diffraction Studies and was performed at the Resource Center for Studies in Surface Science. The TG analysis was carried out in the Center of Thermal Analysis and Calorimetry.

Conflicts of Interest

The authors declare no conflict of interest. The funders played no role in the design of the study; in the collection, analysis or interpretation of data; in the writing of the manuscript or in the decision to publish the results.

References

  1. Brussino, P.; Banús, E.D.; Ulla, M.A.; Bortolozzi, J.P. NiO-based ceramic structured catalysts for ethylene production: Substrates and active sites. Catal. Today 2022, 383, 84–92. [Google Scholar] [CrossRef]
  2. Coomb, D. EVP, Global Olefins & Polyolefins, Chemical Intensity Conference. 2016. Available online: https://www.lyondellbasell.com/globalassets/investors/events/2016/160314-lyb-gs-chemical-intensity-conference-final.pdf (accessed on 15 March 2016).
  3. Gao, Y.; Jiang, J.; Meng, Y.; Yan, F.; Aihemaiti, A. A review of recent developments in hydrogen production via biogas dry reforming. Energy Convers. Manag. 2018, 171, 133–155. [Google Scholar] [CrossRef]
  4. Takanabe, K.; Nagaoka, K.; Nariai, K.; Aika, K. Influence of reduction temperature on the catalytic behavior of Co/TiO2 catalysts for CH4/CO2 reforming and its relation with titania bulk crystal structure. J. Catal. 2005, 230, 75–85. [Google Scholar] [CrossRef]
  5. Fan, X.; Liu, Z.; Zhu, Y.A.; Tong, G.; Zhang, J.; Engelbrekt, C.; Ulstrup, J.; Zhu, K.; Zhou, X. Tuning the composition of metastable CoxNiyMg100−x−y(OH)(OCH3) nanoplates for optimizing robust methane dry reforming catalyst. J. Catal. 2015, 330, 106–119. [Google Scholar] [CrossRef]
  6. Seiyama, T. Total Oxidation of Hydrocarbons on Perovskite Oxides. Catal. Rev. 1992, 34, 281–300. [Google Scholar] [CrossRef]
  7. Voorhoeve, R.J.H.; Johnson, D.W., Jr.; Remeika, J.P.; Gallagher, P.K. Perovskite oxides: Materials science in catalysis. Science 1977, 195, 827–833. [Google Scholar] [CrossRef] [PubMed]
  8. Singh, U.G.; Li, J.; Bennett, J.W.; Rappe, A.M.; Seshadri, R.; Scott, S.L. A Pd-doped perovskite catalyst, BaCe1−xPdxO3, for CO oxidation. J. Catal. 2007, 249, 349–358. [Google Scholar] [CrossRef]
  9. Bashan, V.; Ust, Y. Perovskite catalysts for methane combustion: Applications, design, effects for reactivity and partial oxidation. Int. J. Energy Res. 2019, 43, 7755–7789. [Google Scholar] [CrossRef]
  10. Saada, Y.; Álvarez-Serrano, I.; López, M.L.; Hidouria, M. Structural and dielectric characterization of new lead-free perovskites in the (SrTiO3)–(BiFeO3) system. Ceram. Int. 2016, 42, 8962–8973. [Google Scholar] [CrossRef]
  11. Giannakas, A.E.; Leontiou, A.A.; Ladavos, A.K.; Pomonis, P.J. Characterization and catalytic investigation of NO + CO reaction on perovskites of the general formula LaxM1−xFeO3 (M = Sr and/or Ce) prepared via a reverse micelles microemulsion route. Appl. Catal. A Gen. 2006, 309, 254–262. [Google Scholar] [CrossRef]
  12. Kulkarni, A.S.; Jayaram, R.V. Liquid phase catalytic transfer hydrogenation of aromatic nitro compounds on perovskites prepared by microwave irradiation. Appl. Catal. A Gen. 2003, 252, 225–230. [Google Scholar] [CrossRef]
  13. Goldwasser, M.R.; Dorantes, V.E.; Pérez-Zurit, M.J.; Sojo, P.R.; Cubeiro, M.L.; Pietri, E.; González-Jiménez, F.; NgLee, Y.; Moronta, D. Modified iron perovskites as catalysts precursors for the conversion of syngas to low molecular weight alkenes. J. Mol. Catal. A Chem. 2003, 193, 227–236. [Google Scholar] [CrossRef]
  14. Silva-Santana, M.C.; daSilva, C.A.; Barrozo, P.; Plaza, E.J.R.; Valladares, L.d.; Moreno, N.O. Magnetocaloric and magnetic properties of SmFe0.5Mn0.5O3 complex perovskite. J. Magn. Magn. Mater. 2016, 401, 612–617. [Google Scholar] [CrossRef]
  15. Kesić, Ž.; Lukić, I.; Zdujić, M.; Jovalekić, Č.; Veljković, V.; Skala, D. Assessment of CaTiO3, CaMnO3, CaZrO3 and Ca2Fe2O5 perovskites as heterogeneous base catalysts for biodiesel synthesis. Fuel Process. Technol. 2016, 143, 162–168. [Google Scholar] [CrossRef]
  16. Forni, L.; Rossetti, I. Catalytic combustion of hydrocarbons over perovskites. Appl. Catal. B Environ. 2002, 38, 29–37. [Google Scholar] [CrossRef]
  17. Galasso, F.S. Structure of Perovskite-Type Compounds; A Volume in International Series of Monographs in Solid State Physics; Elsevier Inc.: Amsterdam, The Netherlands, 1969; pp. 3–49. [Google Scholar] [CrossRef]
  18. Moure, C.; Peña, O. Recent advances in perovskites: Processing and properties. Prog. Solid State Chem. 2015, 43, 123–148. [Google Scholar] [CrossRef]
  19. Royer, S.; Duprez, D.; Can, F.; Courtois, X.; Batiot-Dupeyrat, C.; Laassiri, S.; Alamdari, H. Perovskites as Substitutes of Noble Metals for Heterogeneous Catalysis: Dream or Reality. Chem. Rev. 2014, 114, 10292. [Google Scholar] [CrossRef]
  20. Lopatin, S.I.; Zvereva, I.A.; Chislova, I.V. Vaporization and Thermodynamic Properties of GdFeO3 and GdCoO3 Complex Oxides. Russ. J. Gen. Chem. 2020, 90, 1495–1500. [Google Scholar] [CrossRef]
  21. Peña, M.A.; Fierro, J.L.G. Chemical Structures and Performance of Perovskite Oxides. Chem. Rev. 2001, 101, 1981–2018. [Google Scholar] [CrossRef]
  22. Oh, J.H.; Kwon, B.W.; Cho, J.; Lee, C.H.; Kim, M.K.; Choi, S.H.; Yoon, S.P.; Han, J.; Nam, S.W.; Kim, J.Y.; et al. Importance of Exsolution in Transition-Metal (Co, Rh, and Ir)-Doped LaCrO3 Perovskite Catalysts for Boosting Dry Reforming of CH4 Using CO2 for Hydrogen Production. Ind. Eng. Chem. Res. 2019, 58, 6385–6393. [Google Scholar] [CrossRef]
  23. Wang, H.; Dong, X.; Zhao, T.; Yu, H.; Li, M. Dry reforming of methane over bimetallic Ni-Co catalyst prepared from La(CoxNi1−x)0.5Fe0.5O3 perovskite precursor: Catalytic activity and coking resistance. Appl. Catal. B Environ. 2019, 245, 302–313. [Google Scholar] [CrossRef]
  24. Gao, S.; Liu, N.; Liu, J.; Chen, W.; Liang, X.; Yuan, Y. Synthesis of higher alcohols by CO hydrogenation over catalysts derived from LaCo1−xMnxO3 perovskites: Effect of the partial substitution of Co by Mn. Fuel 2020, 261, 116415. [Google Scholar] [CrossRef]
  25. Zhu, Q.; Cheng, H.; Zou, X.; Lu, X.; Xu, Q.; Zhou, Z. Synthesis, characterization, and catalytic performance of La0.6Sr0.4NixCo1−xO3 perovskite catalysts in dry reforming of coke oven gas. Chin. J. Catal. 2015, 36, 915–924. [Google Scholar] [CrossRef]
  26. Bedel, L.; Roger, A.C.; Estournes, C.; Kiennemann, A. Co0 from partial reduction of La(Co,Fe)O3 perovskites for Fischer-Tropsch synthesis. Catal. Today. 2003, 85, 207–218. [Google Scholar] [CrossRef]
  27. Bedel, L.; Roger, A.C.; Rehspringer, J.L.; Zimmermann, Y.; Kiennemann, A. La(1-y)Co0.4Fe0.6O3-δ perovskite oxides as catalysts for Fischer-Tropsch synthesis. J. Catal. 2005, 235, 279–294. [Google Scholar] [CrossRef]
  28. Bedel, L.; Roger, A.C.; Rehspringer, J.L.; Kiennemann, A. Structure-controlled La-Co-Fe perovskite precursors for higher C2-C4 olefins selectivity in Fischer-Tropsch synthesis. Stud. Surf. Sci. Catal. 2004, 147, 319–324. [Google Scholar] [CrossRef]
  29. Sheshko, T.F.; Markova, E.B.; Sharaeva, A.A.; Kryuchkova, T.A.; Zvereva, I.A.; Chislova, I.V.; Yafarova, L.V. Carbon Monoxide Hydrogenation over Gd(Fe/Mn)O3 Perovskite-Type Catalysts. Pet. Chem. 2019, 59, 1307–1313. [Google Scholar] [CrossRef]
  30. Escalona, N.; Fuentealba, S.; Pecchi, G. Fischer-Tropsch synthesis over LaFe1−xCoxO3 perovskites from a simulated biosyngas feed. Appl. Catal. A Gen. 2010, 381, 253–260. [Google Scholar] [CrossRef]
  31. Ding, M.; Yang, Y.; Wu, B.; Li, Y.; Wang, T.; Ma, L. Study on reduction and carburization behaviors of iron phases for iron-based Fischer-Tropsch synthesis catalyst. Energy Proc. 2014, 61, 2267–2270. [Google Scholar] [CrossRef] [Green Version]
  32. Sheshko, T.F.; Kryuchkova, T.A.; Yafarova, L.V.; Borodina, E.M.; Serov, Y.M.; Zvereva, I.A.; Cherednichenko, A.G. Gd-Co-Fe perovskite mixed oxides as catalysts for dry reforming of methane. Sustain. Chem. Pharm. 2022, 30, 100897. [Google Scholar] [CrossRef]
  33. Yafarova, L.V.; Mamontov, G.V.; Chislova, I.V.; Silyukov, O.I.; Zvereva, I.A. The Effect of Transition Metal Substitution in the Perovskite-Type Oxides on the Physicochemical Properties and the Catalytic Performance in Diesel Soot Oxidation. Catalysts 2021, 11, 1256. [Google Scholar] [CrossRef]
  34. Sheshko, T.F.; Serov, Y.M.; Kryuchkova, T.A.; Khayrullina, I.A.; Chislova, I.V.; Yafarova, L.V.; Zvereva, I.A. Study of effect of preparation method and composition on the catalytic properties of complex oxides (Gd,Sr)n+1FenO3n+1 for dry reforming of methane. Nanotechnol. Russ. 2017, 12, 174–184. [Google Scholar] [CrossRef]
  35. Wu, Y.; Li, L.; Chu, B.; Yi, Y.; Qin, Z.; Fan, M.; Qin, Q.; He, H.; Zhang, L.; Dong, L.; et al. Catalytic reduction of NO by CO over B-site partially substituted LaM0.25Co0.75O3 (M = Cu, Mn, Fe) perovskite oxide catalysts: The correlation between physicochemical properties and catalytic performance. Appl. Catal. A Gen. 2018, 568, 43–53. [Google Scholar] [CrossRef]
  36. Merino, N.A.; Barbero, B.P.; Eloy, P.; Cadús, L.E. La1−xCaxCoO3 perovskite-type oxides: Identification of the surface oxygen species by XPS. Appl. Surf. Sci. 2006, 253, 1489–1493. [Google Scholar] [CrossRef]
  37. Wang, Y.; Ren, J.; Wang, Y.; Zhang, F.; Liu, X.; Guo, Y.; Lu, G. Nanocasted synthesis of mesoporous LaCoO3 perovskite with extremely high surface area and excellent activity in methane combustion. J. Phys. Chem. C 2008, 112, 15293–15298. [Google Scholar] [CrossRef]
  38. Sun, J.; Zhao, Z.; Li, Y.; Yu, X.; Zhao, L.; Li, J.; Wei, Y.; Liu, J. Synthesis and catalytic performance of macroporous La1−xCexCoO3 perovskite oxide catalysts with high oxygen mobility for catalytic combustion of soot. J. Rare Earths Chin. Soc. Rare Earths 2020, 38, 584–593. [Google Scholar] [CrossRef]
  39. Kryuchkova, T.A.; Kost’, V.V.; Sheshko, T.F.; Chislova, I.V.; Yafarova, L.V.; Zvereva, I.A. Effect of Cobalt in GdFeO3 Catalyst Systems on Their Activity in the Dry Reforming of Methane to Synthesis Gas. Pet. Chem. 2020, 60, 609–615. [Google Scholar] [CrossRef]
  40. Chislova, I.V.; Matveeva, A.A.; Volkova, A.V.; Zvereva, I.A. Sol-gel synthesis of nanostructured perovskite-like gadolinium ferrites. Glass Phys. Chem. 2011, 37, 653–660. [Google Scholar] [CrossRef]
  41. Yafarova, L.V.; Chislova, I.V.; Zvereva, I.A.; Kryuchkova, T.A.; Kost, V.V.; Sheshko, T.F. Sol–gel synthesis and investigation of catalysts on the basis of perovskite-type oxides GdMO3 (M = Fe, Co). J. Sol. Gel Sci. Technol. 2019, 92, 264–272. [Google Scholar] [CrossRef]
  42. Osman, M.E.; Maximov, V.V.; Dorokhov, V.S.; Mukhin, V.M.; Sheshko, T.F.; Kooyman, P.J.; Kogan, V.M. Carbon-Supported KCoMoS2 for Alcohol Synthesis from Synthesis Gas. Catalysts 2021, 11, 1321. [Google Scholar] [CrossRef]
Figure 1. Dependence of structural parameters a, b and c on composition GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1).
Figure 1. Dependence of structural parameters a, b and c on composition GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1).
Catalysts 13 00008 g001
Figure 2. Micrographs of the surface of GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1).
Figure 2. Micrographs of the surface of GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1).
Catalysts 13 00008 g002
Figure 3. Survey XPS spectrum for compounds in the composition GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) and identification of the main photoelectron lines.
Figure 3. Survey XPS spectrum for compounds in the composition GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) and identification of the main photoelectron lines.
Catalysts 13 00008 g003aCatalysts 13 00008 g003b
Figure 4. O1s spectra for compounds GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) [33].
Figure 4. O1s spectra for compounds GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) [33].
Catalysts 13 00008 g004
Figure 5. Temperature dependences of CO conversions when the reaction is carried out at a CO:H2 = 1:2 ratio over GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8, 1).
Figure 5. Temperature dependences of CO conversions when the reaction is carried out at a CO:H2 = 1:2 ratio over GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8, 1).
Catalysts 13 00008 g005
Figure 6. Temperature dependences of the CO2 formation rates over GdCoxFe1−xO3 complex oxides (CO:H2 ratio = 1:2; GHSV = 8700 h−1).
Figure 6. Temperature dependences of the CO2 formation rates over GdCoxFe1−xO3 complex oxides (CO:H2 ratio = 1:2; GHSV = 8700 h−1).
Catalysts 13 00008 g006
Figure 7. Temperature dependences of (a) methane, (b) ethylene and (c) propylene formation rates on GdCoxFe1−xO3 catalysts.
Figure 7. Temperature dependences of (a) methane, (b) ethylene and (c) propylene formation rates on GdCoxFe1−xO3 catalysts.
Catalysts 13 00008 g007aCatalysts 13 00008 g007b
Figure 8. Chain growth factors αi at 673 K for i = 2, 3, 4 over GdCoxFe1−xO3 catalysts.
Figure 8. Chain growth factors αi at 673 K for i = 2, 3, 4 over GdCoxFe1−xO3 catalysts.
Catalysts 13 00008 g008
Figure 9. Temperature dependences of selectivity for light olefins over GdCoxFe1−xO3 catalysts.
Figure 9. Temperature dependences of selectivity for light olefins over GdCoxFe1−xO3 catalysts.
Catalysts 13 00008 g009
Figure 10. Mass loss curves for GdFeO3, GdCo0.5Fe0.5O3 and GdCoO3 samples.
Figure 10. Mass loss curves for GdFeO3, GdCo0.5Fe0.5O3 and GdCoO3 samples.
Catalysts 13 00008 g010
Table 1. Results of studying GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) surface composition by X-ray spectral, microanalysis, XRD crystallite size and SBET.
Table 1. Results of studying GdCoxFe1−xO3 (x = 0; 0.2; 0.5; 0.8; 1) surface composition by X-ray spectral, microanalysis, XRD crystallite size and SBET.
CompoundXRD Crystallite Size, nmSBET
(m2/g)
Content (at %)Fe/Co
GdFeCoO
GdFeO353.49.918.5319.15-62.32-
GdCo0.2Fe0.8O347.33.317.5816.354.6761.403.5
GdCo0.5Fe0.5O350.52.717.438.779.4964.310.92
GdCo0.8Fe0.2O365.23.116.843.9617.2161.980.23
GdCoO359.92.316.02-17.8066.18-
Table 2. Results of studying the surface composition and state of atoms by XPS [33].
Table 2. Results of studying the surface composition and state of atoms by XPS [33].
CompoundFeCoRatio
Co2+/
∑ Con+
Fe2+/
Fe2+ + Fe3
Os/Ol
GdFeO3+3--0.641.52
GdCo0.2Fe0.8O3+2; +3+2; +30.510.430.59
GdCo0.5Fe0.5O3+2; +3+2; +30.500.580.81
GdCo0.8Fe0.2O3+2; +3+2; +30.130.580.82
GdCoO3-+2; +30.47-0.83
Table 3. Resulting atomic state, CO conversion, selectivity for olefins, experimental activation energies and pre-exponential multipliers for GdCoxFe1−xO3 catalysts.
Table 3. Resulting atomic state, CO conversion, selectivity for olefins, experimental activation energies and pre-exponential multipliers for GdCoxFe1−xO3 catalysts.
CompoundGdFeO3GdCo0.2Fe0.8O3GdCo0.5Fe0.5O3GdCo0.8Fe0.2O3GdCoO3
Co2+/∑ Con+-0.520.500.130.47
Fe2+/Fe2+ + Fe3+0.640.430.580.58-
Os/Ol1.520.590.810.820.83
α(CO), %8044639389
S(CnH2n), %27.138.426.42.40.55
Ea(CH4), kJ/mol707799177/252159
ln K0(CH4)5.75.611.329.6/46.625.1
R20.960.970.990.99/0.990.99
Ea(C2H4), kJ/mol11288115/186112/24499/226
ln K0(C2H4)18.56.712.9/28.215.0/43.111.5/37.1
R20.970.970.96/0.990.96/0.990.98/0.99
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Borodina, E.M.; Yafarova, L.V.; Kryuchkova, T.A.; Sheshko, T.F.; Cherednichenko, A.G.; Zvereva, I.A. Influences of Co-Content on the Physico-Chemical and Catalytic Properties of Perovskite GdCoxFe1−xO3 in CO Hydrogenation. Catalysts 2023, 13, 8. https://doi.org/10.3390/catal13010008

AMA Style

Borodina EM, Yafarova LV, Kryuchkova TA, Sheshko TF, Cherednichenko AG, Zvereva IA. Influences of Co-Content on the Physico-Chemical and Catalytic Properties of Perovskite GdCoxFe1−xO3 in CO Hydrogenation. Catalysts. 2023; 13(1):8. https://doi.org/10.3390/catal13010008

Chicago/Turabian Style

Borodina, Elizaveta M., Liliya V. Yafarova, Tatiana A. Kryuchkova, Tatiana F. Sheshko, Alexander G. Cherednichenko, and Irina A. Zvereva. 2023. "Influences of Co-Content on the Physico-Chemical and Catalytic Properties of Perovskite GdCoxFe1−xO3 in CO Hydrogenation" Catalysts 13, no. 1: 8. https://doi.org/10.3390/catal13010008

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop