Next Article in Journal
Shock Initiation and Propagation of Detonation in ANFO
Previous Article in Journal
Operational Issues of Contemporary Distribution Systems: A Review on Recent and Emerging Concerns
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ceria–Zirconia-Supported Ruthenium Catalysts for Hydrogen Production by Ammonia Decomposition

by
Vadim A. Borisov
1,
Zaliya A. Fedorova
2,
Victor L. Temerev
1,
Mikhail V. Trenikhin
1,
Dmitry A. Svintsitskiy
2,
Ivan V. Muromtsev
1,
Alexey B. Arbuzov
1,
Alexey B. Shigarov
2,
Pavel V. Snytnikov
2,* and
Dmitry A. Shlyapin
1
1
Center of New Chemical Technologies BIC, Boreskov Institute of Catalysis SB RAS, Neftezavodskaya St., 54, 644040 Omsk, Russia
2
Boreskov Institute of Catalysis SB RAS, Lavrentiev Ave. 5, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Energies 2023, 16(4), 1743; https://doi.org/10.3390/en16041743
Submission received: 30 December 2022 / Revised: 25 January 2023 / Accepted: 3 February 2023 / Published: 9 February 2023

Abstract

:
Commercial cerium–zirconium oxide supports (Ce0.5Zr0.5O2, Ce0.75Zr0.25O2, and Ce0.4Zr0.5Y0.05La0.05O2) were used to prepare Ru/CeZrOx catalysts. According to the XRD and IR spectroscopy data, the supports consist of ceria-based substitutional solid solutions. The specific surface areas of supports and catalysts are similar and range from 71–89 m2/g. As shown by TEM and XRD methods, the size of support particles equals 6–11 nm. According to the TEM data, the size of ruthenium particles does not exceed 1.3 nm. The catalyst activity in the ammonia decomposition process was studied. The Ru/Ce0.75Zr0.25O2 catalyst at temperature 500 °C and GHSV 120,000 h−1 demonstrated the highest hydrogen productivity of 53.3 mmol H2/(gcat·min) and compares well with the best results reported in the literature. The kinetics of ammonia decomposition reaction were calculated using the Temkin–Pyzhov exponential expression. The developed mathematical model well described the experimental data. The studied catalysts demonstrated high activity for the ammonia decomposition reaction.

1. Introduction

Recently, hydrogen has attracted particular attention as an environmentally conscious energy carrier which is zero-carbon and provides high energy conversion efficiency when used in fuel cells [1]. However, low volumetric energy density, low boiling point, and the hard-to-handle storage and transportation of hydrogen are still limiting factors for its wide application as a fuel [2]. In 2020, a demonstration project was implemented for hydrogen transportation from Brunei to Japan, using a liquid organic hydrogen compound (LOHC), namely, toluene-methylcyclohexane (TOL-MCH) [3,4]. Currently, the gravimetric storage capacity of the most promising LOHCs does not exceed 7 wt.% hydrogen; in particular, TOL-MCH contains only 6.1 wt.% hydrogen, which is much lower compared to ammonia [5]. Another downside of the LOHC technology is the need to transport the dehydrogenated organic compound back to the hydrogenation plant. The ammonia-based technology is free of this problem; the released nitrogen can either be used for local needs, or simply emitted to the atmosphere. Exactly ammonia seems to be the most attractive hydrogen carrier owing to its environmental safety, carbon-neutrality, and high hydrogen content (17.8 wt.%, or 121 kgH2/m3 at 10 bar); this material is easy to handle and can be readily liquefied, transported, and converted to hydrogen [6].
Hydrogen production from the decomposition of ammonia has been the subject of significant research in recent years. The ammonia decomposition reaction is endothermic and a high operating temperature is required to perform this process at high conversions. Thus, in recent years, significant progress has been made in the research of ammonia decomposition catalysts, and various catalyst systems based on Fe [7,8,9], Co [9,10], and Ni [8,9,11,12,13] have been developed. It is known that ruthenium-based systems are the best catalysts for low-temperature decomposition of ammonia. A large number of studies have been devoted to the research and development of high-performance ruthenium catalysts for the process of ammonia decomposition to produce COx-free hydrogen [4,5,6,14,15,16,17,18,19,20]. Ruthenium catalysts supported by various materials—MgO [9,19,21,22], carbon materials [16,17,18,23,24], Al2O3 [6,18,25,26], ZrO2 [14,27,28,29] zeolites [30], SiO2 [31,32,33], etc.—have been studied in detail. Particular attention is focused on the oxides of rare earth metals La2O3 [27,34,35,36], Pr6O11 [27,37,38,39] and, especially, CeO2 [18,25,26,35,36], which have specific redox and structural characteristics that enhance the catalyst performance in the reaction of ammonia decomposition. For example, Hu et al. found that CeO2 was the best support for the Ru catalyst compared to CNTs and Al2O3, particularly at low temperatures [18]. This result is attributed to the strong metal support interaction and the electronic modification of Ru by ceria.
In order to improve the catalyst activity, thermal stability, and dispersion of the active component, ceria is often doped with oxides of other rare earth metals, for example, ZrO2 [29,40,41,42]. As reported by several research groups, the introduction of trivalent rare earth ions (La3+, Nd3+, Pr3+, Y3+ and Sm3+) into the lattice of CeO2–ZrO2 solid solutions could further improve their reducibility and thermal stability [43], and markedly improve the interaction between the active component and support.
Most often, the synthesis of ruthenium catalysts is performed by impregnating the support with the active component precursor, followed by thermal treatment [18,24,26,27,31,36]. In recent years, attempts have been made to optimize the synthesis procedure in order to obtain the most efficient catalysts containing highly dispersed ruthenium particles and possessing advanced activity in the process of ammonia decomposition.
The present work is devoted to the preparation, physicochemical characterization, and studies of catalytic properties of the ceria–zirconia-supported ruthenium catalysts in the process of hydrogen production by ammonia decomposition. A kinetic model of the process of catalytic decomposition of ammonia was reliably developed describing the experimental results.

2. Experimental

2.1. Synthesis of Ruthenium Catalysts

Catalyst samples containing ~2–3 wt.% ruthenium were prepared. As supports, we chose Ce0.5Zr0.5O2, Ce0.75Zr0.25O2, and Ce0.4Zr0.5Y0.05La0.05O2, to be used as constituent parts of catalysts for automotive exhaust gases afterburning. The active component precursor was the [Ru(NH3)nClm]Clp (n = 5–6; m = 0–1; p = 1–2) complex prepared according to the procedure described in [44]. The synthesized ruthenium–ammonia complex was loaded on the supports by impregnation followed by drying in air at 120 °C for 3 h and reduction in H2 flow (60 mL/min) at 450 °C for 4 h.

2.2. Physicochemical Methods for the Supports and Catalysts Characterization

X-ray diffraction analysis (XRD) of the samples was performed using a D8 Advance diffractometer (Bruker, Germany, CuKα radiation, λ = 0.15418 nm). The diffraction patterns were obtained by scanning in the range 2θ = 10–80 °C, step size 0.05 ° (2θ), and scan speed 2 s/step. Based on the XRD data, the positions of intensity maxima, interplanar spacings dhkl, and full width at half maximum (FWHM) were calculated; the qualitative phase composition of the samples, the unit cell parameters, and dispersion (the coherent scattering regions (CSR) size) were determined.
The low temperature nitrogen adsorption (LTNA) method was used to determine the specific surface area of the supports and catalysts using an ASAP 2400 instrument.
X-ray photoelectron spectroscopy (XPS) measurements were performed using an ES300 spectrometer (KRATOS Analytical, USA) with MgKα (hν = 1253.6 eV). The power of the X-ray source in all cases was 78 W. The instrument was calibrated by the position of the maximum of the Au4f7/2 and Cu2p3/2 lines at 84.0 and 932.7 eV for bulk metallic gold and copper, respectively. The spectra were calibrated by the position of the Ce 3D U‴ peak. In the case of cerium oxide, the position of the U‴ peak was assumed at 916.7 eV [45]. The relative element composition ratio of the sample was determined from perspective peak areas using empirical atomic sensitivity factors (ASF) after recording narrow regions of the corresponding atomic layers [46].
Transmission electron microscopy (TEM) studies were performed using a JEM-2100 transmission electron microscope (JEOL, Japan) at grating resolution 0.14 nm and accelerating voltage 200 kV.

CO Chemisorption

The dispersion of ruthenium in the prepared catalysts was determined by the method of pulsed chemisorption of CO with a ChemBET Pulsar TPR/TPD instrument (Quantachrome Instruments, USA). CO chemisorption was performed after the preliminary reduction of samples in a 10% H2/Ar mixture at 400 °C for 10 min followed by purging with helium for cooling the sample to room temperature. The pulses of 10 vol.% CO/He mixture at regular time intervals was fed into a flow of inert carrier gas (helium). After oxidation in O2 at room temperature, the ruthenium dispersion was redetermined. The values of dispersion and particle sizes were calculated by assuming linear CO chemisorption on ruthenium (the stoichiometric ratio, CO:Ru = 1:1) [47].
The IR spectra of the supports and catalysts were recorded using an IRPresige-21 instrument (Shimadzu, Japan), wavenumber range 350–7900 cm–1, spectral resolution 4 cm–1, and scans accumulation number 50. To record IR spectra, a small amount of the sample was mixed with KBr powder and compressed into a tablet.
Determination of the catalyst activity in the ammonia decomposition process was performed using a fixed-bed flow reactor-based setup (Figure 1).
The reaction of ammonia decomposition was performed in a 12 cm3 glass reactor, at atmospheric pressure, in the temperature range 400–500 °C, and at ammonia flow rates of 20, 60, and 100 mL/min. For the studies, catalyst weights were used: 0.0579 g Ru/Ce0.5Zr0.5O2 05, 0.0511 g Ru/Ce0.75Zr0.25O2, and 0.0525 g Ru/Ce0.4Zr0.5Y0.05La0.05O2. A thermocouple in the reactor was placed in the middle of the catalyst bed. The reaction products were analyzed using a TSVET-500M chromatograph (Russia) with a thermal conductivity detector, the carrier gas—hydrogen; the temperature of the column packed with the Haeyesep C sorbent (which enabled NH3 and N2 separation) was 70 °C. Continuous sampling was provided by a six-way computer-controlled valve. The kinetics were studied in the temperature range 350–510 °C in increments of 20 °C. The catalyst weight was 20.3 mg, so that the degree of decomposition of ammonia did not exceed 40%, even at 510 °C. The best catalyst (Ru/Ce0.75Zr0.25O2) was used to conduct durability tests (Figure S5), with its thermal stability checked for 2 days. Moreover, two such experiments were made—the first with ammonia from a mixture of NH4Cl and NaOH, and the second with pure ammonia. For the first experiment, 0.1077 g was used, and for the second, 0.1058 g. However, for the control in the first experiment, pure ammonia was supplied in the morning and evening.

3. Results and Discussion

According to the LTNA data, the specific surface area of the supports amounted to 71–81 m2/g (Table 1). After ruthenium loading, the specific surface area changed insignificantly (Table 2).
According to the XRD data, the support material is a mixed oxide with a fluorite-type structure (Figure S1). Table 1 presents the values of the unit cell parameter (UCP) and CSR size for mixed oxides Ce1−xZrxO2. In each case, the UCP value is lower than that of unmodified CeO2 (a = 5.411Å) [48], which proves the inclusion of zirconium cations (of a smaller radius) into the CeO2 structure with the formation of a substitutional solid solution. It is known that the introduction of Y3+ and La3+ ions into the lattice of the CeO2–ZrO2 solid solution has a positive effect on the redox properties and oxygen capacity owing to the formation of oxygen vacancies in the initial oxide [43,49]. The XRD patterns of the supports (Figure S1) contain no individual peaks of pure oxides (Ce, Zr, Y, La) because the oxides of rare earth elements completely dissolve in the cerium oxide lattice with the formation of a solid solution.
Table 2 presents the UCP values of the mixed oxide Ce1−xZrxO2, and the CSR sizes for ruthenium catalysts. The XRD diffraction patterns of all catalysts (Figure S2) have no peaks corresponding to ruthenium, which means that ruthenium particles present in a highly dispersed state. Since the measurement accuracy of the UCP and CSR values varies, it seems reasonable to conclude that the catalyst preparation procedure had no effect on the support characteristics.
According to the TEM data, ruthenium particles in the Ru/Ce0.5Zr0.5O2 and Ru/Ce0.4Zr0.5Y0.05La0.05O2 catalysts are uniformly distributed over the support surface (Figure 2a–c). However, the TEM image of the Ru/Ce0.75Zr0.25O2 catalyst showed no ruthenium particles on the support surface. Most likely, ruthenium particles in the latter sample occur in a highly dispersed state, and therefore possess insufficient contrast for visualization in a TEM image. Furthermore, ruthenium can occur in a form of oxidized Ru4+ species which are also invisible in a TEM image due to low contrast ability. Figure 2 also presents the size distribution of the support particles for all catalysts and that of ruthenium particles for Ru/Ce0.5Zr0.5O2 and Ru/Ce0.4Zr0.5Y0.05La0.05O2. It can be seen that the average particle size of the support was 6.4 nm for Ru/Ce0.4Zr0.5Y0.05La0.05O2, 7.5 nm for Ru/Ce0.5Zr0.5O2, and 8.2 nm for Ru/Ce0.75Zr0.25O2 (Table 2). The average particle size of ruthenium was 1.3 nm in Ru/Ce0.5Zr0.5O2 and 1.2 nm in Ru/Ce0.4Zr0.5Y0.05La0.05O2 (Table 2). It should be noted that ruthenium particles of similar size were also obtained in other supports. For example, as reported in [50], a catalyst with ruthenium particles of size 1–2 nm was obtained by impregnating an La2O2CO3–Al2O3 support. In our earlier works, we obtained 3–5 nm ruthenium particles on carbon supports by the impregnation method [44]; with decreasing ruthenium content, the particle size decreased to 2–3 nm [51].
In the present work, the dispersion of ruthenium particles was also determined by the method of CO pulse chemisorptions and compared with the respective TEM data (Table 2).
It can be seen that the high dispersion of ruthenium is confirmed by both methods. In contrast to TEM, the CO chemisorption method allowed the determining ruthenium particle size in Ru/Ce0.75Zr0.25O2, which equaled 2.6 nm. This observation is associated with the reduction of ruthenium from the oxidized state under experimental conditions.
The supports and catalysts were studied by the IR spectroscopic method. The IR spectra of the supports demonstrate intensive absorption bands in the range of 373–380 cm−1, which is characteristic for vibrations of the F1u symmetry in cubic CeO2 [50,51] (Figure 3, spectra 1–3). In addition, broad low-intensive absorption bands were observed in the range of 560–620 and 720–740 cm–1, which are attributed to vibrations of the F1u symmetry in cubic ZrO2 [52]. Thus, in the studied supports, both CeO2 and ZrO2 maintain their cubic structure, in good agreement with the XRD data (Figure S2, Table 1).
After depositing the Ru active component, the vibrational frequency of CeO2 shifts to the high-frequency region by 5–22 cm−1 (Figure 3, spectra 4–6). A similar increase in the vibrational frequency was observed earlier with the introduction of 7 mol.% Ho3+ ions in CeO2 [52].
It should be noted that it is hardly feasible to reveal the structure of RuOx compounds by IR spectroscopy, because some of the compounds (for example, RuO2) have no absorption bands in the 250–1000 cm−1 range. Therefore, auxiliary physicochemical methods should be applied to elucidate the structure of ruthenium compounds.
The XPS method was used to determine the elemental composition of the catalyst surface, and the element chemical states. According to the XPS data, the surface of all catalysts contained the following elements: Ru, Ce, Zr, C, and O (Figure S3).
The XPS spectra of all samples were decomposed in the C 1s and Ru 3D spectral regions and the contribution from the C 1s signal was subtracted. The obtained spectra were a combination of the main Ru 3d5/2 doublet peak at 281.6 ± 0.1 eV and the satellite peak at Eb(Ru 3d5/2) ~ 283 eV (Figure 4). The maximum of the Ru 3p3/2 peak in all cases was close to 463.4 eV. Based on the literature data, the observed spectral characteristics (Eb(Ru 3d5/2) ~ 281.6 eV and Eb(Ru 3p3/2) ~ 463.4 эB) were predominantly assigned to the Ru4+ state [53,54]. The satellite peak at Eb(Ru 3d5/2) ~282.5–283 eV was ascribed to the excitation of plasmons from RuO2-type structures [55], or to the effect of ruthenium cations final state in the rutile structure [56]. Furthermore, the presence of the satellite peak and its intensity can be determined by the crystallization degree of the RuO2 particles, and by the hydration degree of their surface [53,57].
The presence of ruthenium on the catalyst surface, predominantly in an oxidized state, as proved by the XPS data, is most likely explained by the formation of an oxide film resulting from the contact of pre-reduced active component particles with atmospheric air.
For all the catalysts, the Ce 3D spectral regions were decomposed into peaks corresponding to the Ce3+ and Ce4+ states. It was found that in all samples, the content of the Ce3+ species on the support surface was almost the same and equaled 21 ± 3%; cerium in the Ce4+ state dominated, in good agreement with the literature data [58,59].
The Zr 3D spectra of all samples were correctly described by a single doublet component with the Zr 3d5/2 binding energy at ~ 181.9 eV, corresponding to a Zr4+ state similar to that in the ZrO2 composition (Figure 5) [46,60].
In the Ru/Ce0.4Zr0.5Y0.05La0.05O2 catalyst, the elements La and Y were not revealed by the XPS method (Figure S4), but both elements were detected in the catalyst by chemical analysis (Table S2). This observation is most likely explained by their localization in the bulk and low content on the support surface.
The catalysts were tested in the process of ammonia decomposition. The results of catalytic tests conducted at various ammonia flow rates and temperatures (400, 440, and 500 °C) are presented in Table S2. Figure 6 illustrates the results of catalyst testing at an ammonia flow rate of 100 mL/min. It can be seen that with increasing temperature, the conversion of ammonia over all catalysts increases significantly. At 500 °C, the ammonia conversion over catalysts Ru/Ce0.5Zr0.5O2 and Ru/Ce0.75Zr0.25O2 is practically the same, but the latter catalyst demonstrates much higher hydrogen productivity (53.3 mmol H2/min·gcat). In addition, compared to other catalysts, this sample showed a sharper increase in activity with increasing temperature, i.e., it has the maximum activation energy. Note also that the specific activity of this catalyst exceeded the activity of the known catalyst synthesized in [61].
The kinetics of ammonia decomposition over the Ru/Ce0.5Zr0.5O2 catalyst were studied as a function of the hydrogen content in the reaction mixture formed by mixing the reagent flows to provide a total flow rate of the mixture of 100 mL/min. The hydrogen concentration was varied in the range of 5–90 vol.% by changing its flow rate from 5 mL/min to 90 mL/min at a reaction temperature of 490 °C. At a reaction temperature of 410 °C, the hydrogen flow rate was varied from 10 mL/min to 90 mL/min. At a constant ammonia content of 10 vol.%, this means that the H2/NH3 ratio was 0.5 ÷ 9.0. To maintain a constant flow rate of the reaction mixture, helium was introduced into the system (up to 100 mL/min). The results of catalytic tests, linearized using the Van ’t Hoff differential method, are shown in Figure 7a.
As Figure 7a shows, when the hydrogen content in the reaction mixture increases from 5 to 55 vol.% (corresponding lnC(H2) interval is 1.5 to 4), the reaction rate decreases linearly (at 490 °C). The calculation shows that in this H2 concentration interval, the apparent reaction order with respect to hydrogen is −0.27. When the hydrogen content in the reaction mixture varies from 65 to 90 vol.% (corresponding lnC(H2) interval is 4–4.5), the reaction rate decreases much more sharply, and the apparent reaction order with respect to hydrogen approaches −2.59. Obviously, Ru/Ce0.5Zr0.5O2, similarly to other ruthenium catalysts, is susceptible to the hydrogen inhibition effect described in [62]. In the literature, the effect of hydrogen inhibition is explained by hydrogen displacement from the catalyst surface of nitrogen-containing transitional forms (N, NH, NH2), which have a lower adsorption heat than that of hydrogen. The concentration of these nitrogen species is an important factor for both decomposition and synthesis of ammonia; thus, the effect of hydrogen inhibition is significant for both forward and reverse reactions. In the range of 50–100 vol.% hydrogen concentrations (Figure 7a, 410 °C, purple dots), the apparent reaction order with respect to hydrogen is −0.53.
The effect of ammonia content in the reaction mixture on the catalytic properties of the Ru/Ce0.5Zr0.5O2 was studied in the range of 10–100 vol.%. inlet NH3 concentrations. The results of catalytic tests, linearized using the Van ’t Hoff differential method, are shown in Figure 7b. Indeed, in the 10–50 vol.% NH3 concentration range (corresponding lnC(NH3) interval is 2–4 (490 °C)), the reaction rate increases proportionally to increasing ammonia content. As the ammonia concentration increases within 50–100 vol.% (490 °C) (corresponding lnC(NH3) range is 4 to 4.5), the apparent reaction order with respect to ammonia decreases to 0.25. The observed point, in which the character of the dependence alters (the inflection point), corresponds to the reaction mixture composition of 50 vol.% NH3, 50 vol.% He.
At the reaction temperature of 410 °C, no change in the apparent reaction order is indicated; we can instead discuss the common order of the reaction in the studied NH3 concentration interval, equal to 0.13. The reason for this catalyst behavior at 410 °C is most likely related to the rearrangement of the catalyst active centers owing to the influence of the support. This reminds that cerium–zirconium mixed oxides are phases characterized by high oxygen lability, which can transit to an oxygen-lean state in a reducing medium, such as that containing ammonia and hydrogen. The oxygen deficit most likely promotes the formation of a partial positive charge in the support phase that affects the character of its electronic interaction with ruthenium particles. Probably, depending on the temperature, the surface oxygen deficit varies, thus altering the electronic state of the active component in contact with the reaction medium.
In general, it can be assumed that the observed dependence results from the competitive adsorption of ammonia and hydrogen on the ruthenium surface. The concentration ratio of the released H2 and undecomposed NH3 in the gas mixture exceeds 0.8 at the inlet reaction mixture composition of 50 vol.% NH3 and 50 vol.% He (inflection point at 490 °C). At high concentrations of ammonia in a mixture with helium, its conversion does not exceed 30%, and the NH3-to-hydrogen concentration ratio falls below 0.5. As the NH3 concentration decreases, its conversion increases and the hydrogen concentration also increases. As a result, the competition between hydrogen and ammonia for the active centers of the catalyst becomes more profound.
The observed dependencies are characteristic for the competitive adsorption of reagents on the catalyst surface. A similar situation was observed in our recent study [63] on acetylene hydrogenation to ethylene over the Pd-based catalyst: the competitive adsorption of acetylene and hydrogen was proved by the inflection point on the curve of the catalyst activity depending on the acetylene concentration in the reaction mixture.
Temperature is an important factor that significantly affects both the catalyst activity and the reaction product distribution. Figure 8 presents the results of studying the influence of temperature increasing from 350 to 510 °C with a step of 20 °C on the catalytic performance of the properties of Ru/Ce0.5Zr0.5O2.
Obviously, after linearization, all points fit on a straight line. The apparent activation energy Ea calculated by the Arrhenius equation equals ~90 kJ/mol.
The activation energy value complies with that reported in the literature for ruthenium catalysts on oxide supports [62,64,65,66].

Simulation of Experimental Data on Ammonia Decomposition over the Ru/Ce0.5Zr0.5O2 Catalyst Using Kinetic Data Reported in the Literature

Strictly speaking, comparative analysis of the activity of various catalysts in terms of specific productivity in a flow reactor is correct only at a low conversion of the initial reagent that ensures an insignificant concentration gradient along the reactor length. Otherwise, if the apparent reaction order is positive with respect to the inlet reagent (ammonia in our case) and negative with respect to the product (hydrogen), the apparent productivity for any catalyst will noticeably decrease with increasing reagent (ammonia) conversion. To overcome this problem, it seems reasonable to compare different catalysts at similar conversions (see Table 2), but this approach is hardly feasible experimentally, especially when comparison with the literature data is needed.
An alternative approach to compare the catalysts’ activity is based on mathematical modeling of the reactor in which the tests were performed, using global kinetics. Note that for the studied catalyst, the kinetics themselves may still be unknown. However, we can try to model the experimental data on the ammonia conversion for a particular catalyst using the kinetic data obtained by other authors for other catalysts (with a similar active component and support) and, thus, compare this catalyst in terms of volumetric activity with other catalysts with already known kinetic data.
This approach was applied to analyze the activity of one of the catalysts studied in this work, namely, Ru/Ce0.5Zr0.5O2. Unfortunately, there are very few published kinetic data appropriate for simulating a fixed-bed flow reactor for catalysts of this composition. We were able to find only two relatively close examples, which are briefly considered below.
The authors of [64] used a flow differential fixed-bed flow reactor and a 3%LiOH/1%Ru/α-Al2O3 catalyst in the temperature range of 350–475 °C and obtained a Temkin–Pyzhev-like power-law expression with reversibility term:
r N H 3 = k 0   exp ( E A R   T )   ( P N H 3 )   0.24 ( P H 2 )   0.54   ( 1 [ ( P H 2 )   1.5   ( P N 2 )   0.5 K e q   P N H 3 ] 2 )
with E A = 82.6 kJ/mol and k 0 = 36.18 mol(NH3)*kPa0.3/gcat/s.
In [24], using a fixed-bed flow reactor for x%Ni-y%Ru/CeO2 bimetallic catalysts at x = 1.5–10 wt.% and y = 0.3–2 wt.%, the authors developed the Langmuir–Hinshelwood– Hougen–Watson (LHHW) kinetic model and obtained the best agreement between the calculated and experimental results under the assumption that the rate-limiting stage is the dehydrogenation of the first hydrogen atom (the remaining stages were considered as quasi-stationary):
N H 3 ( a d s ) + [ * ] N H 2 ( a d s ) + H ( a d s )
The respective kinetic equation is as follows:
r N H 3 = k 0   exp ( E A R   T )   K N H 3 C N H 3 ( 1 + K N H 3 C N H 3 + C H 2 K H 2   ) 2
For the 5%Ni-1%Ru/CeO2 catalyst, the authors of [24] assumed the following kinetic parameters: E A = 124 kJ/mol, k 0 = 3.9 × 106 mol(NH3)/gcat/s, K N H 3 = 1.1 m3/mol, and K H 2 = 4.7 × 10−2 mol/m3.
Calculations were performed according to an isothermal model of a plug–flow reactor using kinetic expressions and parameters (1) and (3) under the operating conditions of experiments with the Ru/Ce0.5Zr0.5O2 catalyst (layer bulk density ~1000 kg/m3): normal pressure, inlet composition—NH3 100%, ammonia flow rate—100 mL(gas)/min, and GHSV—296,000 h−1. For the calculations, we used the in-home-made program code. The calculations were performed in the temperature range of 350–500 °C with a step of 10 °C. A correction multiplier for the pre-exponential factor was selected in order to provide the coincidence of the calculated ammonia conversion at 450 °C with the experimental value of 10.5% for both kinetic Equations (1) and (3).
Figure 9 compares the calculated and experimental results for ammonia conversion depending on the temperature. Curve 1 was obtained by 10-fold increasing the pre-exponential factor for kinetic Equation (1). Curve 2 was obtained by 3-fold increasing the kinetic Equation (3). That is, at 450 °C, our Ru/Ce0,5Zr0,5O2 catalyst is an order of magnitude more active than the 3%LiOH/1%Ru/α-Al2O3 catalyst (which, in turn, is close in activity to unpromoted Ru/α-Al2O3 catalysts [64]), and three times more active than the 5%Ni-1%Ru/CeO2 catalyst. Taking into account that the ruthenium content in the latter catalyst is 2–3 times lower, it seems reasonable to conclude that our Ru/Ce0.5Zr0.5O2 catalyst has almost as high activity per 1 g Ru as that of the 5%Ni-1%Ru/CeO2 [24], and exceeds it 3-fold in terms of volumetric activity.
A comparison table of our catalysts with other catalysts, as well as correlations between the physical and catalytic properties of our catalysts are given in additional materials. Additional materials also contain references [67,68,69,70,71,72,73,74,75,76,77,78,79] and explanations to the comparison table and correlations.

4. Conclusions

  • The IR spectroscopy and XRD data proved the formation of solid CeO2–ZrO2 solutions in the supported ruthenium catalysts. With increasing zirconium content, the crystallite size decreases and equals 8.2 nm for Ru/Ce0.75Zr0.25O2, 7.5 nm for Ru/Ce0.5Zr0.5O2, and 6.4 nm for Ru/Ce0.4Zr0.5Y0.05La0.05O2. The synthesized catalysts are characterized by high dispersion of the active component. According to the TEM data, the ruthenium particle size in Ru/Ce0.5Zr0.5O2 and Ru/Ce0.4Zr0.5Y0.05La0.05O2 was 1.3 and 1.2 nm, respectively.
  • According to the XPS data, ruthenium on the catalyst surface occurs predominantly in the Ru4+ state, most probably due to its oxidation at contact with air. Cerium in all samples presents predominantly in the Ce4+ state.
  • Ruthenium catalysts were tested in the process of ammonia decomposition. The best activity showed the catalyst in support with the highest content of cerium (Ru/Ce0.75Zr0.25O2). It provided a hydrogen productivity of 53.3 mmol H2/min•gcat at 500 °C and retained high activity during 48 h testing in the ammonia decomposition process.
  • The kinetics of NH3 decomposition were calculated using the Temkin–Pyzhov exponential expression. The mathematic model describes well with the experimental data.

5. Notations

  • W is the specific rate of reaction, mmol(H2)/(gcat min);
  • r N H 3 is the specific rate of reaction, mol(NH3)/gcat/s;
  • C N H 3 and C H 2 are the ammonia and hydrogen concentrations, respectively, mol/m3;
  • P N H 3 , P H 2 and P N 2 are the ammonia, hydrogen, and nitrogen partial pressure bars, respectively;
  • k 0 is the pre-exponential factor;
  • K e q is the temperature-dependent equilibrium constant;
  • E A is the activation energy;
  • K N H 3 is the ammonia adsorption constant;
  • K H 2 is the hydrogen desorption constant.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en16041743/s1.

Author Contributions

Conceptualization, V.A.B., Z.A.F., V.L.T. and P.V.S.; Methodology, M.V.T., D.A.S. (Dmitry A. Svintsitskiy), I.V.M., A.B.A. and D.A.S. (Dmitry A. Shlyapin); Validation, V.A.B. and D.A.S. (Dmitry A. Shlyapin); Formal analysis, M.V.T., D.A.S. (Dmitry A. Svintsitskiy), I.V.M., A.B.S. and D.A.S. (Dmitry A. Shlyapin); Investigation, V.A.B., Z.A.F., V.L.T., M.V.T., D.A.S. (Dmitry A. Svintsitskiy), I.V.M. and A.B.A.; Data curation, V.A.B., Z.A.F. and A.B.S.; Writing—original draft, V.A.B., Z.A.F., V.L.T., M.V.T. and D.A.S. (Dmitry A. Shlyapin); Visualization, V.A.B., Z.A.F., A.B.A. and A.B.S.; Supervision, V.A.B.; Project administration, P.V.S.. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Higher Education of the Russian Federation within the governmental order for the Boreskov lnstitute of Catalysis (project AAAA-A21-121011390009-1).

Data Availability Statement

Not applicable.

Acknowledgments

Physicochemical studies were conducted using the equipment of the Omsk Regional Center of Collective Usage SB RAS and Shared-Use Center “National Center for the Study of Catalysts” at the BIC.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Qureshi, F.; Yusuf, M.; Kamyab, H.; Vo, D.-V.N.; Chelliapan, S.; Joo, S.-W.; Vasseghian, Y. Latest eco-friendly avenues on hydrogen production towards a circular bioeconomy: Currents challenges, innovative insights, and future perspectives. Renew. Sustain. Energy Rev. 2022, 168, 112916. [Google Scholar] [CrossRef]
  2. Qureshi, F.; Yusuf, M.; Kamyab, H.; Zaidi, S.; Khalil, M.J.; Arham Khan, M.; Azad Alam, M.; Masood, F.; Bazli, L.; Chelliapan, S.; et al. Current trends in hydrogen production, storage and applications in India: A review. Sustain. Energy Technol. Assess. 2022, 53, 102677. [Google Scholar] [CrossRef]
  3. Raab, M.; Maier, S.; Dietrich, R.-U. Comparative techno-economic assessment of a large-scale hydrogen transport via liquid transport media. Int. J. Hydrogen Energy 2021, 46, 11956–11968. [Google Scholar] [CrossRef]
  4. Chiyoda. Mitsubishi link with Singaporean firms for sustainable hydrogen economy. Fuel Cells Bull. 2020, 5, 14. [Google Scholar] [CrossRef]
  5. Abdin, Z.; Tang, C.; Liu, Y.; Catchpole, K. Large-scale stationary hydrogen storage via liquid organic hydrogen carriers. IScience 2021, 24, 102966. [Google Scholar] [CrossRef] [PubMed]
  6. Nagaoka, K.; Eboshi, T.; Takeishi, Y.; Tasaki, R.; Honda, K.; Imamura, K.; Sato, K. Carbon-free H2 production from ammonia triggered at room temperature with an acidic RuO2 /γ-Al2 O3 catalyst. Sci. Adv. 2017, 3, e1602747. [Google Scholar] [CrossRef] [PubMed]
  7. Hu, Z.-P.; Chen, L.; Chen, C.; Yuan, Z.-Y. Fe/ZSM-5 catalysts for ammonia decomposition to COx-free hydrogen: Effect of SiO2/Al2O3 ratio. Mol. Catal. 2018, 455, 14–22. [Google Scholar] [CrossRef]
  8. Simonsen, S.B.; Chakraborty, D.; Chorkendorff, I.; Dahl, S. Alloyed Ni-Fe nanoparticles as catalysts for NH3 decomposition. Appl. Catal. A Gen. 2012, 447–448, 22–31. [Google Scholar] [CrossRef]
  9. Hu, X.-C.; Wang, W.-W.; Jin, Z.; Wang, X.; Si, R.; Jia, C.-J. Transition metal nanoparticles supported La-promoted MgO as catalysts for hydrogen production via catalytic decomposition of ammonia. J. Energy Chem. 2019, 38, 41–49. [Google Scholar] [CrossRef]
  10. Lendzion-Bielun, Z.; Narkiewicz, U.; Arabczyk, W. Cobalt-based Catalysts for Ammonia Decomposition. Materials 2013, 6, 2400–2409. [Google Scholar] [CrossRef] [Green Version]
  11. Henpraserttae, S.; Charojrochkul, S.; Lawtrakul, L.; Toochinda, P. Ni-based Catalysts for Hydrogen Production from Ammonia Decomposition: Effect of Dopants and Urine Application. Chem. Select 2018, 3, 11842–11850. [Google Scholar] [CrossRef]
  12. Yu, Y.; Gan, Y.-M.; Huang, C.; Lu, Z.-H.; Wang, X.; Zhang, R.; Feng, G. Ni/La2O3 and Ni/MgO–La2O3 catalysts for the decomposition of NH3 into hydrogen. Int. J. Hydrog. Energy 2020, 45, 16528–16539. [Google Scholar] [CrossRef]
  13. Nakamura, I.; Fujitani, T. Role of metal oxide supports in NH3 decomposition over Ni catalysts. Appl. Catal. A Gen. 2016, 524, 45–49. [Google Scholar] [CrossRef]
  14. Lorenzut, B.; Montini, T.; Pavel, C.C.; Comotti, M.; Vizza, F.; Bianchini, C.; Fornasiero, P. Embedded Ru@ZrO2 Catalysts for H2 Production by Ammonia Decomposition. ChemCatChem 2010, 2, 1096–1106. [Google Scholar] [CrossRef]
  15. Borisov, V.A.; Iost, K.N.; Temerev, V.L.; Simunin, M.M.; Leont’eva, N.N.; Mikhlin, Y.L.; Volochaev, M.N.; Shlyapin, D.A. Ammonia decomposition Ru catalysts supported on alumina nanofibers for hydrogen generation. Mater. Lett. 2022, 306, 130842. [Google Scholar] [CrossRef]
  16. Mukherjee, S.; Devaguptapu, S.V.; Sviripa, A.; Lund, C.R.F.; Wu, G. Low-temperature ammonia decomposition catalysts for hydrogen generation. Appl. Catal. B Env. 2018, 226, 162–181. [Google Scholar] [CrossRef]
  17. Fujitani, T.; Nakamura, I.; Hashiguchi, Y.; Kanazawa, S.; Takahashi, A. Effect of Catalyst Preparation Method on Ammonia Decomposition Reaction over Ru/MgO Catalyst. Bull. Chem. Soc. Jpn. 2020, 93, 1186–1192. [Google Scholar] [CrossRef]
  18. Lucentini, I.; García Colli, G.; Luzi, C.D.; Serrano, I.; Martínez, O.M.; Llorca, J. Catalytic ammonia decomposition over Ni-Ru supported on CeO2 for hydrogen production: Effect of metal loading and kinetic analysis. Appl. Catal. B Environ. 2021, 286, 119896. [Google Scholar] [CrossRef]
  19. Ju, X.; Liu, L.; Zhang, X.; Feng, J.; He, T.; Chen, P. Highly Efficient Ru/MgO Catalyst with Surface-Enriched Basic Sites for Production of Hydrogen from Ammonia Decomposition. ChemCatChem 2019, 11, 4161–4170. [Google Scholar] [CrossRef]
  20. Ju, X.; Liu, L.; Yu, P.; Guo, J.; Zhang, X.; He, T.; Wu, G.; Chen, P. Mesoporous Ru/MgO prepared by a deposition-precipitation method as highly active catalyst for producing COx-free hydrogen from ammonia decomposition. Appl. Catal. B Environ. 2017, 211, 167–175. [Google Scholar] [CrossRef]
  21. Hu, Z.; Mahin, J.; Datta, S.; Bell, T.E.; Torrente-Murciano, L. Ru-Based Catalysts for H2 Production from Ammonia: Effect of 1D Support. Top. Catal. 2019, 62, 1169–1177. [Google Scholar] [CrossRef]
  22. Zheng, W.; Zhang, J.; Zhu, B.; Blume, R.; Zhang, Y.; Schlichte, K.; Schlögl, R.; Schüth, F.; Su, D.S. Structure-Function Correlations for Ru/CNT in the Catalytic Decomposition of Ammonia. ChemSusChem 2010, 3, 226–230. [Google Scholar] [CrossRef] [PubMed]
  23. Hill, A.K.; Torrente-Murciano, L. Low temperature H2 production from ammonia using ruthenium-based catalysts: Synergetic effect of promoter and support. Appl. Catal. B Env. 2015, 172–173, 129–135. [Google Scholar] [CrossRef]
  24. Lucentini, I.; Casanovas, A.; Llorca, J. Catalytic ammonia decomposition for hydrogen production on Ni, Ru and Ni-Ru supported on CeO2. Int. J. Hydrogen Energy 2019, 44, 12693–12707. [Google Scholar] [CrossRef]
  25. Hu, X.C.; Fu, X.P.; Wang, W.W.; Wang, X.; Wu, K.; Si, R.; Ma, C.; Jia, C.J.; Yan, C.H. Ceria-supported ruthenium clusters transforming from isolated single atoms for hydrogen production via decomposition of ammonia. Appl. Catal. B Environ. 2020, 268, 118424. [Google Scholar] [CrossRef]
  26. Wang, S.J.; Yin, S.F.; Li, L.; Xu, B.Q.; Ng, C.F.; Au, C.T. Investigation on modification of Ru/CNTs catalyst for the generation of COx-free hydrogen from ammonia. Appl. Catal. B Environ. 2004, 52, 287–299. [Google Scholar] [CrossRef]
  27. Im, Y.; Muroyama, H.; Matsui, T.; Eguchi, K. Investigation on catalytic performance and desorption behaviors of ruthenium catalysts supported on rare-earth oxides for NH3 decomposition. Int. J. Hydrogen Energy 2022, 47, 32543–32551. [Google Scholar] [CrossRef]
  28. Wang, Z.; Qu, Y.; Shen, X.; Cai, Z. Ruthenium catalyst supported on Ba modified ZrO2 for ammonia decomposition to COx-free hydrogen. Int. J. Hydrogen Energy 2019, 44, 7300–7307. [Google Scholar] [CrossRef]
  29. Miyamoto, M.; Hamajima, A.; Oumi, Y.; Uemiya, S. Effect of basicity of metal doped ZrO2 supports on hydrogen production reactions. Int. J. Hydrogen Energy 2018, 43, 730–738. [Google Scholar] [CrossRef]
  30. Cha, J.; Lee, T.; Lee, Y.-J.; Jeong, H.; Jo, Y.S.; Kim, Y.; Nam, S.W.; Han, J.; Lee, K.B.; Yoon, C.W.; et al. Highly monodisperse sub-nanometer and nanometer Ru particles confined in alkali-exchanged zeolite Y for ammonia decomposition. Appl. Catal. B Env. 2021, 283, 119627. [Google Scholar] [CrossRef]
  31. Wang, F.; Deng, L.; Wu, Z.; Ji, K.; Chen, Q.; Jiang, X. The dispersed SiO2 microspheres supported Ru catalyst with enhanced activity for ammonia decomposition. Int. J. Hydrog. Energy 2021, 46, 20815–20824. [Google Scholar] [CrossRef]
  32. Li, Y.; Yao, L.; Song, Y.; Liu, S.; Zhao, J.; Ji, W.; Au, C.-T. Core–shell structured microcapsular-like Ru@SiO2 reactor for efficient generation of COx-free hydrogen through ammonia decomposition. Chem. Commun. 2010, 46, 5298. [Google Scholar] [CrossRef] [PubMed]
  33. Varisli, D.; Elverisli, E.E. Synthesizing hydrogen from ammonia over Ru incorporated SiO2 type nanocomposite catalysts. Int. J. Hydrog. Energy 2014, 39, 10399–10408. [Google Scholar] [CrossRef]
  34. Huang, C.; Yu, Y.; Yang, J.; Yan, Y.; Wang, D.; Hu, F.; Wang, X.; Zhang, R.; Feng, G. Ru/La2O3 catalyst for ammonia decomposition to hydrogen. Appl. Surf. Sci. 2019, 476, 928–936. [Google Scholar] [CrossRef]
  35. Ogura, Y.; Sato, K.; Miyahara, S.; Kawano, Y.; Toriyama, T.; Yamamoto, T.; Matsumura, S.; Hosokawa, S.; Nagaoka, K. Efficient ammonia synthesis over a Ru/La0.5Ce0.5O1.75 catalyst pre-reduced at high temperature. Chem. Sci. 2018, 9, 2230–2237. [Google Scholar] [CrossRef] [PubMed]
  36. Han, W.; Li, Z.; Liu, H. La2Ce2O7 supported ruthenium as a robust catalyst for ammonia synthesis. J. Rare Earths 2019, 37, 492–499. [Google Scholar] [CrossRef]
  37. Sato, K.; Imamura, K.; Kawano, Y.; Miyahara, S.; Yamamoto, T.; Matsumura, S.; Nagaoka, K. A low-crystalline ruthenium nano-layer supported on praseodymium oxide as an active catalyst for ammonia synthesis. Chem. Sci. 2017, 8, 674–679. [Google Scholar] [CrossRef]
  38. Nagaoka, K.; Honda, K.; Ibuki, M.; Sato, K.; Takita, Y. Highly Active Cs2/Ru/Pr6O11as a Catalyst for Ammonia Decomposition. Chem. Lett. 2010, 39, 918–919. [Google Scholar] [CrossRef]
  39. Nagaoka, K.; Eboshi, T.; Abe, N.; Miyahara, S.I.; Honda, K.; Sato, K. Influence of basic dopants on the activity of Ru/Pr6O11 for hydrogen production by ammonia decomposition. Int. J. Hydrog. Energy 2014, 3, 20731–20735. [Google Scholar] [CrossRef]
  40. Carbajal-Ramos, I.A.; Gomez, M.F.; Condó, A.M.; Bengió, S.; Andrade-Gamboa, J.J.; Abello, M.C.; Gennari, F.C. Catalytic behavior of Ru supported on Ce0.8Zr0.2O2 for hydrogen production. Appl. Catal. B Env. 2016, 181, 58–70. [Google Scholar] [CrossRef]
  41. Atribak, I.; Guillén-Hurtado, N.; Bueno-López, A.; García-García, A. Influence of the physico-chemical properties of CeO2–ZrO2 mixed oxides on the catalytic oxidation of NO to NO2. Appl. Surf. Sci. 2010, 256, 7706–7712. [Google Scholar] [CrossRef]
  42. Nakonieczny, D.; Paszenda, Z.; Drewniak, S.; Radko, T.; Lis, M. sZrO2-CeO2 ceramic powders obtained from a sol-gel process using acetylacetone as a chelating agent for potential application in prosthetic dentistry. Acta Bioeng. Biomech. 2016, 18, 53–60. [Google Scholar] [PubMed]
  43. Zhao, B.; Wang, Q.; Li, G.; Zhou, R. Effect of rare earth (La, Nd, Pr, Sm and Y) on the performance of Pd/Ce0.67Zr0.33MO2−δ three-way catalysts. J. Environ. Chem. Eng. 2013, 1, 534–543. [Google Scholar] [CrossRef]
  44. Borisov, V.A.; Iost, K.N.; Petrunin, D.A.; Temerev, V.L.; Muromtsev, I.V.; Arbuzov, A.B.; Trenikhin, M.V.; Gulyaeva, T.I.; Smirnova, N.S.; Shlyapin, D.A.; et al. Effect of the Modifier on the Catalytic Properties and Thermal Stability of Ru–Cs(Ba)/Sibunit Catalyst for Ammonia Decomposition. Kinet. Catal. 2019, 60, 372–379. [Google Scholar] [CrossRef]
  45. Slavinskaya, E.M.; Gulyaev, R.V.; Stonkus, O.A.; Zadesenets, A.V.; Plyusnin, P.E.; Shubin, Y.V.; Korenev, S.V.; Ivanova, A.S.; Zaikovskii, V.I.; Danilova, I.G.; et al. Low-temperature oxidation of carbon monoxide on Pd(Pt)/CeO2 catalysts prepared from complex salts. Kinet. Catal. 2011, 52, 282–295. [Google Scholar] [CrossRef]
  46. Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D.; Chastain, J. Handbook of X-Ray Photoelectron Spectroscopy; Physical Electronics Division, Perkin-Elmer Corporation: Eden Prairie, MN, USA, 1992. [Google Scholar]
  47. Maroto-Valiente, A.; Cerro-Alarcón, M.; Guerrero-Ruiz, A.; Rodríguez-Ramos, I. Effect of the metal precursor on the surface site distribution of Al2O3-supported Ru catalysts: Catalytic effects on the n-butane/H2 test. Appl. Catal. A Gen. 2005, 283, 23–32. [Google Scholar] [CrossRef]
  48. Grier, D.; McCarthy, G. North Dakota State University, ICDD Grant-in-Aid 1991; Powder Diffraction File; JCPDS Data File No. 43-1002, (n.d.); International Center for Diffraction Data: Newtown Square, PA, USA, 1994; pattern 00-043-1002. [Google Scholar]
  49. He, H.; Dai, H.X.; Ng, L.H.; Wong, K.W.; Au, C.T. Pd-, Pt-, and Rh-Loaded Ce0.6Zr0.35Y0.05O2.Three-Way Catalysts: An Investigation on Performance and Redox Properties. J. Catal. 2002, 206, 1–13. [Google Scholar] [CrossRef]
  50. Kim, A.-R.; Cha, J.; Kim, J.S.; Ahn, C.-I.; Kim, Y.; Jeong, H.; Choi, S.H.; Nam, S.W.; Yoon, C.W.; Sohn, H. Hydrogen production from ammonia decomposition over Ru-rich surface on La2O2CO3-Al2O3 catalyst beads. Catal. Today. 2022, in press. [CrossRef]
  51. Borisov, V.A.; Iost, K.N.; Temerev, V.L.; Fedotova, P.A.; Surovikin, Y.V.; Arbuzov, A.B.; Trenikhin, M.V.; Shlyapin, D.A. Effect of high-temperature treatment of on the activity of Ru-Cs(Ba)/Sibunit catalysts in ammonia synthesis and their resistance to methanation. Diam. Relat. Mater. 2020, 108, 107986. [Google Scholar] [CrossRef]
  52. López, E.F.; Escribano, V.S.; Panizza, M.; Carnasciali, M.M.; Busca, G. Vibrational and electronic spectroscopic properties of zirconia powders. J. Mater. Chem. 2001, 11, 1891–1897. [Google Scholar] [CrossRef]
  53. Morgan, D.J. Resolving ruthenium: XPS studies of common ruthenium materials. Surf. Interface Anal. 2015, 47, 1072–1079. [Google Scholar] [CrossRef]
  54. Ernst, M.A.; Sloof, W.G. Unraveling the oxidation of Ru using XPS. Surf. Interface Anal. 2008, 40, 334–337. [Google Scholar] [CrossRef]
  55. Over, H.; Seitsonen, A.P.; Lundgren, E.; Smedh, M.; Andersen, J.N. On the origin of the Ru-3d5=2 satellite feature from RuO2 (110). Surf. Sci. 2002, 504, L196–L200. [Google Scholar] [CrossRef]
  56. Kim, Y.J.; Gao, Y.; Chambers, S.A. Core-level X-ray photoelectron spectra and X-ray photoelectron diffraction of RuO2 (110) grown by molecular beam epitaxy on TiO2 (110). Appl. Surf. Sci. 1997, 120, 250–260. [Google Scholar] [CrossRef]
  57. Foelske, A.; Barbieri, O.; Hahn, M.; Kötz, R. An X-ray photoelectron spectroscopy study of hydrous ruthenium oxide powders with various water contents for supercapacitors. Electrochem. Solid-State Lett. 2006, 9, A268. [Google Scholar] [CrossRef]
  58. Fox, E.B.; Lee, A.F.; Wilson, K.; Song, C. In-situ XPS Study on the Reducibility of Pd-Promoted Cu/CeO2 Catalysts for the Oxygen-assisted Water-gas-shift Reaction. Top. Catal. 2008, 49, 89–96. [Google Scholar] [CrossRef]
  59. Zyryanova, M.M.; Snytnikov, P.V.; Gulyaev, R.V.; Amosov, Y.I.; Boronin, A.I.; Sobyanin, V.A. Performance of Ni/CeO2 catalysts for selective CO methanation in hydrogen-rich gas. Chem. Eng. J. 2014, 238, 189–197. [Google Scholar] [CrossRef]
  60. Wagner, C.D.; Naumkin, A.V.; Kraut-Vass, A.; Allison, J.W.; Powell, C.J.; Rumble, C.J. NIST X-ray Photoelectron Spectroscopy Database, Version 3.5; Natl. Inst. Stand. Technol.: Gaithersbg, MD, USA, 2003. [Google Scholar]
  61. Yin, S.F.; Zhang, Q.H.; Xu, B.Q.; Zhu, W.X.; Ng, C.F.; Au, C.T. Investigation on the catalysis of COx-free hydrogen generation from ammonia. J. Catal. 2004, 224, 384–396. [Google Scholar] [CrossRef]
  62. Takahashi, A.; Fujitani, T. Kinetic Analysis of Decomposition of Ammonia over Nickel and Ruthenium Catalysts. J. Chem. Eng. Japan 2016, 49, 22–28. [Google Scholar] [CrossRef]
  63. Shlyapin, D.A.; Glyzdova, D.V.; Afonasenko, T.N.; Temerev, V.L.; Tsyrul’nikov, P.G. Acetylene Hydrogenation to Ethylene in a Hydrogen-Rich Gaseous Mixture on a Pd/Sibunit Catalyst. Kinet. Catal. 2019, 60, 446–452. [Google Scholar] [CrossRef]
  64. Lamb, K.; Hla, S.S.; Dolan, M. Ammonia decomposition kinetics over LiOH-promoted, A-Al2O3-supported Ru catalyst. Int. J. Hydrog. Energy 2019, 44, 3726–3736. [Google Scholar] [CrossRef]
  65. Sayas, S.; Morlanés, N.; Katikaneni, S.P.; Harale, A.; Solami, B.; Gascon, J. High pressure ammonia decomposition on Ru–K/CaO catalysts. Catal. Sci. Technol. 2020, 10, 5027–5035. [Google Scholar] [CrossRef]
  66. Prasad, V.; Karim, A.M.; Arya, A.; Vlachos, D.G. Assessment of Overall Rate Expressions and Multiscale, Microkinetic Model Uniqueness via Experimental Data Injection: Ammonia Decomposition on Ru/γ-Al2O3 for Hydrogen Production. Ind. Eng. Chem. Res. 2009, 48, 5255–5265. [Google Scholar] [CrossRef]
  67. Chen, J.; Zhu, Z.H.; Wang, S.; Ma, Q.; Rudolph, V.; Lu, G.Q. Effects of nitrogen doping on the structure of carbon nanotubes (CNTs) and activity of Ru/CNTs in ammonia decomposition. Chem. Eng. J. 2010, 156, 404–410. [Google Scholar] [CrossRef]
  68. Kocer, T.; Oztuna, F.E.S.; Kurtoğlu, S.F.; Unal, U.; Uzun, A. Graphene aerogel-supported ruthenium nanoparticles for COx-free hydrogen production from ammonia. Appl. Catal. A Gen. 2021, 610, 117969. [Google Scholar] [CrossRef]
  69. Kurtoğlu, S.F.; Soyer-Uzun, S.; Uzun, A. Utilizing Red Mud Modified by Simple Treatments as a Support to Disperse Ruthenium Provides a High and Stable Performance for COx-Free Hydrogen Production from Ammonia. Catal. Today. 2020, 357, 425–435. [Google Scholar] [CrossRef]
  70. Wang, Z.; Cai, Z.; Wei, Z. Highly Active Ruthenium Catalyst Supported on Barium Hexaaluminate for Ammonia Decomposition to COx-Free Hydrogen. ACS Sustain. Chem. Eng. 2019, 7, 8226–8235. [Google Scholar] [CrossRef]
  71. Hayashi, F.; Toda, Y.; Kanie, Y.; Kitano, M.; Inoue, Y.; Yokoyama, T.; Hara, M.; Hosono, H. Ammonia decomposition by ruthenium nanoparticles loaded on inorganic electride C12A7:e−. Chem. Sci. 2013, 4, 3124–3130. [Google Scholar] [CrossRef]
  72. Le, T.A.; Kim, Y.; Kim, H.W.; Lee, S.-U.; Kim, J.-R.; Kim, T.-W.; Lee, Y.-J.; Chae, H.-J. Ru-supported lanthania-ceria composite as an efficient catalyst for COx-free H2 production from ammonia decomposition. Appl. Catal. B Env. 2021, 285, 119831. [Google Scholar] [CrossRef]
  73. Furusawa, T.; Shirasu, M.; Sugiyama, K.; Sato, T.; Itoh, N.; Suzuki, N. Preparation of Ru/ZrO2 Catalysts by NaBH4 Reduction and Their Catalytic Activity for NH3 Decomposition to Produce H2. Ind. Eng. Chem. Res. 2016, 55, 12742–12749. [Google Scholar] [CrossRef]
  74. Yin, S.-F.; Xu, B.-Q.; Ng, C.-F.; Au, C.-T. Nano Ru/CNTs: A highly active and stable catalyst for the generation of COx-free hydrogen in ammonia decomposition. Appl. Catal. B Env. 2004, 48, 237–241. [Google Scholar] [CrossRef]
  75. Baidya, T.; Dutta, G.; Hegde, M.S.; Waghmare, U.V. Correlating Particle Size and Shape of Supported Ru/γ-Al2O3 Catalysts with NH3 Decomposition Activity. J. Amer. Chem. Soc. 2009, 131, 12230–12239. [Google Scholar] [CrossRef]
  76. Baidya, T.; Dutta, G.; Hegde, M.S.; Waghmare, U.V. Noble metal ionic catalysts: Correlation of increase in CO oxidation activity with increasing effective charge on Pd ion in Pd ion substituted Ce1− x MxO2− δ (M= Ti, Zr and Hf). Dalt. Trans. 2009, 3, 455–464. [Google Scholar] [CrossRef]
  77. Reddy, B.M.; Khan, A. Nanosized CeO2–SiO2, CeO2–TiO2, and CeO2–ZrO2 mixed oxides: Influence of supporting oxide on thermal stability and oxygen storage properties of ceria. Catal. Surv. Asia. 2005, 9, 155–171. [Google Scholar] [CrossRef]
  78. Reddy, B.M.; Khan, A.; Yamada, Y.; Kobayashi, T.; Loridant, S.; Volta, J.C. Structural characterization of CeO2−MO2 (M = Si4+, Ti4+, and Zr4+) mixed oxides by Raman spectroscopy, X-ray photoelectron spectroscopy, and other techniques. J. Phys. Chem. B. 2003, 107, 11475–11484. [Google Scholar] [CrossRef]
  79. Reddy, B.M.; Khan, A.; Yamada, Y.; Kobayashi, T.; Loridant, S.; Volta, J.C. Raman and X-ray photoelectron spectroscopy study of CeO2−ZrO2 and V2O5/CeO2−ZrO2 catalysts. Langmuir 2003, 19, 3025–3030. [Google Scholar] [CrossRef]
Figure 1. A scheme of the ammonia decomposition setup.
Figure 1. A scheme of the ammonia decomposition setup.
Energies 16 01743 g001
Figure 2. TEM images of ruthenium catalysts deposited on ceria–zirconia supports: (a,b) Ru/Ce0.4Zr0.5Y0.005La0.005O2; (c,d) Ru/Ce0.5Zr0.5O2; (e,f) Ru/Ce0.75Zr0.25O2. Size distribution of support particles (blue histogram): (b) Ru/Ce0.4Zr0.5Y0.05La0.05O2; (d) Ru/Ce0.5Zr0.5O2; (f) Ru/Ce0.75Zr0.25O2; size distribution of ruthenium particles (red histogram): (a) Ru/Ce0.4Zr0.5Y0.05La0.05O2, (c) Ru/Ce0.5Zr0.5O2.
Figure 2. TEM images of ruthenium catalysts deposited on ceria–zirconia supports: (a,b) Ru/Ce0.4Zr0.5Y0.005La0.005O2; (c,d) Ru/Ce0.5Zr0.5O2; (e,f) Ru/Ce0.75Zr0.25O2. Size distribution of support particles (blue histogram): (b) Ru/Ce0.4Zr0.5Y0.05La0.05O2; (d) Ru/Ce0.5Zr0.5O2; (f) Ru/Ce0.75Zr0.25O2; size distribution of ruthenium particles (red histogram): (a) Ru/Ce0.4Zr0.5Y0.05La0.05O2, (c) Ru/Ce0.5Zr0.5O2.
Energies 16 01743 g002
Figure 3. IR spectra of supports: 1—Ce0.75Zr0.25O2; 2—Ce0.5Zr0.5O2; 3—Ce0.4Zr0.5Y0.05La0.05O2 and catalysts: 4—Ru/Ce0.75Zr0.25O2; 5—Ru/Ce0.5Zr0.5O2; 6—Ru/Ce0.4Zr0.5Y0.05La0.05O2.
Figure 3. IR spectra of supports: 1—Ce0.75Zr0.25O2; 2—Ce0.5Zr0.5O2; 3—Ce0.4Zr0.5Y0.05La0.05O2 and catalysts: 4—Ru/Ce0.75Zr0.25O2; 5—Ru/Ce0.5Zr0.5O2; 6—Ru/Ce0.4Zr0.5Y0.05La0.05O2.
Energies 16 01743 g003
Figure 4. The decomposed Ru 3D spectra after subtracting the contribution from the C 1s signal for catalysts: (1) Ru/Ce0.5Zr0.5O2; (2) Ru/Ce0.75Zr0.25O2; (3) Ru/Ce0.4Zr0.5Y0.05La0.05O2.
Figure 4. The decomposed Ru 3D spectra after subtracting the contribution from the C 1s signal for catalysts: (1) Ru/Ce0.5Zr0.5O2; (2) Ru/Ce0.75Zr0.25O2; (3) Ru/Ce0.4Zr0.5Y0.05La0.05O2.
Energies 16 01743 g004
Figure 5. The Zr 3D spectra of catalysts: (1) Ru/Ce0.5Zr0.5O2; (2) Ru/Ce0.75Zr0.25O2; (3) Ru/Ce0.4Zr0.5Y0.05La0.05O2.
Figure 5. The Zr 3D spectra of catalysts: (1) Ru/Ce0.5Zr0.5O2; (2) Ru/Ce0.75Zr0.25O2; (3) Ru/Ce0.4Zr0.5Y0.05La0.05O2.
Energies 16 01743 g005
Figure 6. The temperature dependence of ammonia conversion over ruthenium catalysts (P = 1 atm, ammonia flow rate = 100 mL/min, mcat = 50 mg).
Figure 6. The temperature dependence of ammonia conversion over ruthenium catalysts (P = 1 atm, ammonia flow rate = 100 mL/min, mcat = 50 mg).
Energies 16 01743 g006
Figure 7. The logarithmic dependences of the rate of ammonia decomposition over Ru/Ce0.5Zr0.5O2 catalyst on the (a) hydrogen concentration and (b) ammonia concentration in the reaction mixture at 490 and 410 °C.
Figure 7. The logarithmic dependences of the rate of ammonia decomposition over Ru/Ce0.5Zr0.5O2 catalyst on the (a) hydrogen concentration and (b) ammonia concentration in the reaction mixture at 490 and 410 °C.
Energies 16 01743 g007
Figure 8. Linearization for finding the parameters of the Arrhenius equation for the Ru/Ce0.5Zr0.5O2 catalyst.
Figure 8. Linearization for finding the parameters of the Arrhenius equation for the Ru/Ce0.5Zr0.5O2 catalyst.
Energies 16 01743 g008
Figure 9. Ammonia conversion as a function of temperature. Symbols—experimental data for Ru/Ce0.5Zr0.5O2 catalyst. Curve 1—calculation according to the kinetic equation [64] for the 3%LiOH/1%Ru/α-Al2O3 catalyst with a 9.6-fold increase of the pre-exponential factor k 0 . Curve 2—calculation according to the kinetic equation [24] for the 5%Ni-1%Ru/CeO2 catalyst with a 2.8-fold increase of the pre-exponential factor k 0 .
Figure 9. Ammonia conversion as a function of temperature. Symbols—experimental data for Ru/Ce0.5Zr0.5O2 catalyst. Curve 1—calculation according to the kinetic equation [64] for the 3%LiOH/1%Ru/α-Al2O3 catalyst with a 9.6-fold increase of the pre-exponential factor k 0 . Curve 2—calculation according to the kinetic equation [24] for the 5%Ni-1%Ru/CeO2 catalyst with a 2.8-fold increase of the pre-exponential factor k 0 .
Energies 16 01743 g009
Table 1. The main characteristics of mixed oxides Ce1-xZrxO2.
Table 1. The main characteristics of mixed oxides Ce1-xZrxO2.
SupportSBET, m2/gDXRD, nmUCP, nmDTEM, nm
Ce0.75Zr0.25O28710.50.53978.2
Ce0.5Zr0.5O2817.50.52827.4
Ce0.4Zr0.5Y0.05La0.05O2716.80.52616.4
Table 2. The main characteristics of ruthenium catalysts.
Table 2. The main characteristics of ruthenium catalysts.
CatalystRu Content *, wt.%SBET,
m2/g
UCPsup, nmParticle Size, nm
SupportRuthenium
XRDTEMTEMCO Chemisorption
Ru/Ce0.75Zr0.25O22.61890.539510.88.2-2.6
Ru/Ce0.5Zr0.5O23.32800.52857.67.41.32.0
Ru/Ce0.4Zr0.5Y0.05La0.05O22.26840.52566.86.41.21.6
* According to X-ray fluorescence analysis.
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

Borisov, V.A.; Fedorova, Z.A.; Temerev, V.L.; Trenikhin, M.V.; Svintsitskiy, D.A.; Muromtsev, I.V.; Arbuzov, A.B.; Shigarov, A.B.; Snytnikov, P.V.; Shlyapin, D.A. Ceria–Zirconia-Supported Ruthenium Catalysts for Hydrogen Production by Ammonia Decomposition. Energies 2023, 16, 1743. https://doi.org/10.3390/en16041743

AMA Style

Borisov VA, Fedorova ZA, Temerev VL, Trenikhin MV, Svintsitskiy DA, Muromtsev IV, Arbuzov AB, Shigarov AB, Snytnikov PV, Shlyapin DA. Ceria–Zirconia-Supported Ruthenium Catalysts for Hydrogen Production by Ammonia Decomposition. Energies. 2023; 16(4):1743. https://doi.org/10.3390/en16041743

Chicago/Turabian Style

Borisov, Vadim A., Zaliya A. Fedorova, Victor L. Temerev, Mikhail V. Trenikhin, Dmitry A. Svintsitskiy, Ivan V. Muromtsev, Alexey B. Arbuzov, Alexey B. Shigarov, Pavel V. Snytnikov, and Dmitry A. Shlyapin. 2023. "Ceria–Zirconia-Supported Ruthenium Catalysts for Hydrogen Production by Ammonia Decomposition" Energies 16, no. 4: 1743. https://doi.org/10.3390/en16041743

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