Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect
ACS Publications. Most Trusted. Most Cited. Most Read
Evolution of Oxygen Vacancy Sites in Ceria-Based High-Entropy Oxides and Their Role in N2 Activation
My Activity

Figure 1Loading Img
  • Open Access
Energy, Environmental, and Catalysis Applications

Evolution of Oxygen Vacancy Sites in Ceria-Based High-Entropy Oxides and Their Role in N2 Activation
Click to copy article linkArticle link copied!

  • Omer Elmutasim
    Omer Elmutasim
    Mechanical Engineering Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    Center for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
  • Aseel G. Hussien
    Aseel G. Hussien
    Mechanical Engineering Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    Center for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
  • Abhishek Sharan
    Abhishek Sharan
    Center for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    Physics Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
  • Sara AlKhoori
    Sara AlKhoori
    Mechanical Engineering Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    Center for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
  • Michalis A. Vasiliades
    Michalis A. Vasiliades
    Department of Chemistry, Heterogeneous Catalysis Laboratory, University of Cyprus, 1 University Avenue, University Campus, 2109 Nicosia, Cyprus
  • Inas Magdy Abdelrahman Taha
    Inas Magdy Abdelrahman Taha
    Physics Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
  • Seokjin Kim
    Seokjin Kim
    Oxide & Organic Nanomaterials for Energy & Environment (ONE) Laboratory, Advanced Membranes & Porous Materials (AMPM) Center, and KAUST Catalysis Center (KCC), Physical Science & Engineering (PSE), King Abdullah University of Science and Technology (KAUST), Thuwal 23955, Saudi Arabia
    More by Seokjin Kim
  • Messaoud Harfouche
    Messaoud Harfouche
    Synchrotron-Light for Experimental Science and Applications in the Middle East (SESAME), Allan 19252, Jordan
  • Abdul-Hamid Emwas
    Abdul-Hamid Emwas
    Core Laboratories, King Abdullah University of Science and Technology (KAUST), Thuwal 23955-6900, Saudi Arabia
  • Dalaver H. Anjum
    Dalaver H. Anjum
    Center for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    Physics Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
  • Angelos M. Efstathiou
    Angelos M. Efstathiou
    Department of Chemistry, Heterogeneous Catalysis Laboratory, University of Cyprus, 1 University Avenue, University Campus, 2109 Nicosia, Cyprus
  • Cafer T. Yavuz
    Cafer T. Yavuz
    Oxide & Organic Nanomaterials for Energy & Environment (ONE) Laboratory, Advanced Membranes & Porous Materials (AMPM) Center, and KAUST Catalysis Center (KCC), Physical Science & Engineering (PSE), King Abdullah University of Science and Technology (KAUST), Thuwal 23955, Saudi Arabia
  • Nirpendra Singh*
    Nirpendra Singh
    Center for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    Physics Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    *Email: [email protected]
  • Kyriaki Polychronopoulou*
    Kyriaki Polychronopoulou
    Mechanical Engineering Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    Center for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    *Email: [email protected]
Open PDFSupporting Information (1)

ACS Applied Materials & Interfaces

Cite this: ACS Appl. Mater. Interfaces 2024, 16, 18, 23038–23053
Click to copy citationCitation copied!
https://doi.org/10.1021/acsami.3c16521
Published April 29, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

In this work, a relatively new class of materials, rare earth (RE) based high entropy oxides (HEO) are discussed in terms of the evolution of the oxygen vacant sites (Ov) content in their structure as the composition changes from binary to HEO using both experimental and computational tools; the composition of HEO under focus is the CeLaPrSmGdO due to the importance of ceria-related (fluorite) materials to catalysis. To unveil key features of quinary HEO structure, ceria-based binary CePrO and CeLaO compositions as well as SiO2, the latter as representative nonreducible oxide, were used and compared as supports for Ru (6 wt % loading). The role of the Ov in the HEO is highlighted for the ammonia production with particular emphasis on the N2 dissociation step (N2(ads) → Nads) over a HEO; the latter step is considered the rate controlling one in the ammonia production. Density functional theory (DFT) calculations and 18O2 transient isotopic experiments were used to probe the energy of formation, the population, and the easiness of formation for the Ov at 650 and 800 °C, whereas Synchrotron EXAFS, Raman, EPR, and XPS probed the Ce–O chemical environment at different length scales. In particular, it was found that the particular HEO composition eases the Ov formation in bulk, in medium (Raman), and in short (localized) order (EPR); more Ov population was found on the surface of the HEO compared to the binary reference oxide (CePrO). Additionally, HEO gives rise to smaller and less sharp faceted Ru particles, yet in stronger interaction with the HEO support and abundance of Ru–O–Ce entities (Raman and XPS). Ammonia production reaction at 400 °C and in the 10–50 bar range was performed over Ru/HEO, Ru/CePrO, Ru/CeLaO, and Ru/SiO2 catalysts; the Ru/HEO had superior performance at 10 bar compared to the rest of catalysts. The best performing Ru/HEO catalyst was activated under different temperatures (650 vs 800 °C) so to adjust the Ov population with the lower temperature maintaining better performance for the catalyst. DFT calculations showed that the HEO active site for N adsorption involves the Ov site adjacent to the adsorption event.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2024 The Authors. Published by American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

Ammonia is maybe the most important chemical for humankind. The widely used, industrial-scale process for ammonia production is still dependent on the energy-demanding Haber–Bosch (HB) process (temperatures in the 723–773 K and pressures in the 15–30 MPa); the latter is consuming high amounts of energy (nearly above 1% of the global energy) in annual basis. Therefore, the design of processes that are more environmentally friendly and less energy-hungry is of pivotal importance. (1) Shifting the energy demanding HB process to milder conditions would require a catalyst surface where N and H atoms are not bind strongly, while the activation barrier for the N2 to N dissociation is rather low; the latter is inversely dependent to the adsorbed atomic N stability (scaling relations). (2)
Fe and Ru are among the traditional catalysts for ammonia production with Fe being more acceptable for industrial use due to its lower costs. (3,4) The dissociative mechanism involves the adsorption of N2 followed by its dissociation into 2N (atomic nitrogen) before its reaction with H2 and the production of NH3(g). The energy-demanding step is that of N2 activation where the breaking of the N–N stable triple bond is involved; the latter is considered the rate-determining step on Ru and Fe widely used, catalysts. Therefore, a thorough understanding of the active sites and/or the catalyst design criteria involved in the N2 activation is crucial for optimizing the ammonia production rate.
Cerium dioxide (CeO2) is a widely used catalyst component for a span of applications from industrial and automobile exhausts (5) to promote the water gas shift reaction, (6) and hydrogen production reactions (7) as well as solid electrolyte in fuel cells, (8) due to its superior ability to recycle oxygen through redox reactions and high mobility of oxygen ion. The rapid formation and elimination of oxygen vacancy defects plays a crucial role in all these catalytic reactions. Ceria-based high-entropy oxides are expected to withhold great potential for catalytic reactions where the high oxygen mobility is a significant catalyst treatment, (9,10) likely due to the theoretically predicted easiness of formation of the oxygen vacancies. Other properties of HEO that make them attractive candidates for catalysis/electrocatalysis are their high ionic/mixed conductivity due to their high lattice strain (11) as well as the presence of multiple luminescent centers (origin of exceptional optical properties). In particular, the lattice distortion in the HEO contributes to properties such as good transparency, high refractive index, overall enhancing the imaging quality, and expanding viewing angle lens. Moreover, the stabilization of unconventional spin-electronic states can lead to noteworthy magnetic applications. It has been reported that the HEO with spinel structure responds quickly to small magnetic field changes. Additionally, the combination of morphological features with the tunability of the lattice distortion can give rise to distinct values of strength and elastic modulus (mechanical properties). Last but not least, the low thermal conductivity of the high entropy oxides is usually combined with fairly good electrical conductivity reaching the levels of steel and lead. (12−14) In a recent review article by our group, the design criteria of the HEO are explicitly presented and discussed toward tuning the features that dictate their catalysis chemistry. (10) As it has also been eloquently discussed in (15) the multielemental nature of the HEO gives space for tunability of the active sites and/or vacancies location, whereas understanding the defects formation plays a major role in unveiling their functional properties and driving new discoveries. From a thermodynamic perspective, the driving force for defect formation is the configurational entropy increase, without undermining the role of the constituent elements. (9)
In the present study, the evolution of the formation of oxygen vacancy (Ov) defects as a result of stepwise transition from pure ceria (single oxide), to binary, to ternary systems, and eventually to quinary rare earth high entropy oxide (HEO) systems is explored. The CeLaPrSmGdO composition was used as model one. Computational ab initio tools, as well as experimental techniques, such as Raman spectroscopy, X-ray photoelectron spectroscopy (XPS), electron paramagnetic resonance (EPR), and synchrotron X-ray absorption spectroscopy (XAS) along with detailed electron microscopy tools involving HRTEM, STEM-HAADF, RGB analysis, FFT analysis, and isotopic oxygen exchange transient experiments, are employed. Following the investigation of the Ov formation in the HEO composition, a demonstration of the role of Ov in the ammonia thermal production is provided over Ru based HEO supported catalyst with a critical comparison with Ru-based catalysts where the support is either binary reducible oxide (CePrO, CeLaO) or nonreducible one (SiO2). The impact of the support (HEO vs CePrO) on the growth of Ru particles shape and size is critically discussed whereas insights on the N2 dissociation over the Ru/HEO catalyst and the role of the Ov proximity to the N2 dissociation site (HEO active site) are provided.

2. Methods

Click to copy section linkSection link copied!

2.1. Computational Studies

The first-principles calculations are based on density functional theory as this is developed in the Vienna ab initio Simulation package (VASP), (16,17) using generalized gradient approximation (GGA) of Perdew, Burke, and Ernzerhof (18) as the exchange-correlation functional. Projector-augmented-wave potentials are used to account for the valence electrons and ionic core interactions. (19,20) 4f electrons of the rare-earth elements are considered core electrons for structural relaxations, while for computation of the total energy, 4f electrons are considered valence electrons. For computational efficacy, soft pseudopotentials of oxygen and nitrogen are used. The GGA+U method (21) is used to push the unoccupied 4f orbitals of the rare-earth elements away from the Fermi level, and the Hubbard U parameter is set to 5 eV for all the rare-earth elements. Spin-polarized calculations are performed for structural relaxations, while for computations of the total energy spin–orbit interaction is taken into account. A fluorite-type crystal structure (with space group 225) of pure CeO2 is used for computations. The structure of the ternary and quinary high-entropy oxide (HEO) is generated using special quasirandom structures (SQS) (22) based on a 144 and 180-atom supercell respectively on a parent CeO2 prototype structure, using alloy theoretic automated toolkit (ATAT) code. (23) SQS has been successful in describing the electronic and thermodynamic properties of various disordered systems. (24) Energy cutoff of 350 eV is used for plane wave basis set expansion and Γ-centered k-point mesh of 4 × 4 × 4, 3 × 5 × 2, 3 × 2 × 2, and 2 × 1 × 1 is used for pure CeO2, binary, ternary and quinary HEO, respectively. The oxygen vacancies on the (111) surface of these systems are considered, where a vacuum region of 20 Å is used in the supercell containing the surface to avoid spurious interactions with the periodic images. The formation energy (ΔEform) of oxygen vacancy is computed as follows:
ΔEf=Esp:Ovac+12EO2Esp
(1)
where Esp:Ovac is the energy of the supercell with one oxygen vacancy, EO2 is the energy of one oxygen molecule, and Esp is the energy of the supercell without any oxygen vacancy.

2.1.1. Ru on HEO Calculations

The literature revealed that the ruthenium cluster, having a pyramid structure composed of 4 atoms, supported on ceria is an energetically stable structure. (25) Likewise, in this work, a similar Ru4 cluster was deposited atop the oxygen vacancy site on reduced quinary LaPrCeSmGdO10(111) surface (hereafter named the Ru4/HEO surface).

2.2. Experimental Studies

2.2.1. Catalysts Preparation

2.2.1.1. Metal Oxide Synthesis
In this study, the high entropy oxide (CeLaPrSmGdO) support and two reference (reducible) supports (i.e., CeLaO and CePrO) were synthesized via the coprecipitation method. (26) The total moles of the metal nitrate precursors were maintained at 0.023 mol. For the CeLaPrSmGdO (HEO) support, the metal nitrate precursors with equimolar composition were dissolved in 45 mL DI water and placed on an orbital shaker with RPM set to 340 at 65 °C for 15 h. Then, 12.5 mL of ammonium hydroxide was added to the solution while stirring for an extra 2 h. The mixture was dried overnight at 100 °C, then calcined at 900 °C at a rate of 5 °C/min. A similar procedure was followed for the synthesis of reference supports, CeLaO and CePrO, except that the molar ratio of Ce to the metal (i.e., La and Pr) was 4:1. The SiO2 support was provided by Zeofree (commercial) and it was used as received.
2.2.1.2. Ru-Based Catalyst Synthesis
Ruthenium nanoparticles were prepared by chemical reduction of RuCl3·xH2O using NaBH4 as a reducing agent and KOH as a stabilizer (in situ reduction). First, 50 mL of 2-propanol, 50 mL of DI water, and 360 mg of oxide support (HEO, CePrO, CeLaO, SiO2) were mixed together and sonicated for 20 min. Then, 0.047 g of RuCl3·xH2O was added to the mixture and sonicated for 5 more mins. The mixture was stirred on a high magnetic stirrer at 500 rpm for 10 min. Then, a mix of 0.1 M KOH and 0.1 M of NaBH4 (total 50 mL) was slowly added and stirred for an additional 1 h. The final mix was washed 3 times in DI water and then dried overnight at 90 °C. Therefore, the prepared Ru-based catalysts are the: Ru/HEO, Ru/CePrO, Ru/CeLaO, and Ru/SiO2 all with 6 wt % Ru loading.

2.2.2. Characterization of Catalysts

Most of the techniques used in this study are described in previous studies. (27,28) More details about Raman, XPS, EPR, and HRTEM are given in the SI.
2.2.2.1. Synchrotron EXAFS
The X-ray absorption fine structure (XAFS) allows to assess the oxidation state by taking into account the X-ray absorption near edge structure (XANES) segment of the spectra; additionally, it helps to unveil structural characteristics at different length scales- short and medium range- through utilizing the extend X-ray absorption fine structure (EXAFS) part of data; the latter were collected on the BM08-XAFS/XRF beamline (29) at Synchrotron-light for Experimental Science and Applications in the Middle East (SESAME), Jordan for all catalysts of this study following their reduction (10 vol % H2/Ar, 650 °C, 120 min). XAFS spectra were recorded, at room temperature, in step-by-step scanning mode at the Ce LIII-edge (5723 eV). Transmission and fluorescence modes of the signal were collected in XAFS in both ionization chambers and the energy-selective single-element silicon drift detector (SDD) from KETEK GmbH, Germany. For scanning the energy, at the XANES region, with a step size of 0.2 and 0.5 eV for the pre-edge and white line, respectively, a double crystal monochromator equipped with Si(111) crystal was used. Variable energy steps using a fixed k space wavenumber of 0.03 A–1 were utilized for recording the EXAFS region of the spectra. Demeter (30) and WinXAS (31) software packages were used for the collection and processing of the XAFS data. The reference spectra used for XANES are from the databases (32,33) for the CeO2, CePO4, and CeF4.
2.2.2.2. 16O/18O Surface Isotopic Exchange
The solid in powder form (50 mg, dp <106 μm) was introduced into a quartz CSTR microreactor, (34) where the absence of interparticle (external) and intraparticle (internal) concentration gradients was verified. (35) Both catalysts were pretreated as follows: (i) Exp 1: Calcination (5% O2/He) at 600 °C/1 h; (ii) Exp 2: Calcination (5% O2/He) at 600 °C for 1 h, Reduction (5% H2/He) at 600 °C/2 h; (iii) Exp 3: Calcination (5% O2/He) at 600 °C for 1 h, Reduction (5% H2/He) at 800 °C/2 h. Transient Isothermal Oxidation (TIO) was used to investigate the formation of oxygen vacancies as a function of the pretreatment procedure over the two solids. More specifically, following the catalyst pretreatment, the reactor was purged in He for 30 min at 600 °C, followed by a step-gas switch to 1% 16O2/1% Kr/Ar/He (FT = 100 N mL min–1). It should be mentioned that at the end of TIO, steady-state rates of oxygen transfer between gas-phase oxygen and lattice oxygen (OL)/oxygen vacancies (Ov) of the solid are established. TIO experiment was followed by a transient isothermal isotopic exchange with 18O2 experiment (18O2-TIIE) in order to probe the kinetics of 16O/18O surface isotopic exchange and bulk oxygen diffusion (oxygen mobility) under equilibrium or pseudo equilibrium conditions, (36) verified by the sum (1 mol %) of all three isotopologues oxygen gas compositions during the whole transient. It is worth mentioning that the shape and the position of the 16O18O species strongly depend on the bulk oxygen diffusion coefficient (Deff, cm2 s–1) of the solid, but also the amount of oxygen (N16O, mmol g–1) available to be exchanged. (37) Specifically, during 18O2-TIIE the following gas step switch was applied: 1% 16O2/1% Kr/Ar/He → 1% 18O2/Ar/He at 600 °C. During the 18O gas mixture, the transient responses of 16O2, 16O18O, 18O2, and Kr (m/z = 32, 34, 36, and 84, respectively) were continuously monitored via an online mass spectrometer (Balzers, Quadrupole 1–200 amu), and converted into concentration (mol %) by using certified gas mixtures (i.e., 2% 16O2/He, 5% 18O2/He, and the impurity concentration of 16O18O(g) present in the 5% 18O2/ He).
The N16O (mmol g–1) amount of 16O/18O exchanged (oxygen storage capacity, OSC) was estimated via the material balance depicted in eq 2:
N16O[mol g1]=2yf16O2FTW0t(ZO162(t)ZKr(t))dt+FTW0t(yO16O18(t)yO16O18f[1ZKr])dt
(2)
where Zi and yi are the dimensionless response and mole fraction of gaseous species i, respectively.
The dimensionless αg(18)(t) descriptor function illustrates both the surface 16O/18O exchange (initial period under 18O gas), which is crucial from an industrial point of view, and the oxygen diffusion in the bulk (prolonged time on 18O gas), as obtained via eq 3. It should be noted that the lower the value of αg(18)(t), the faster the oxygen diffusion in the bulk of the solid.
αg(18)(t)=yO16O18+2yO1822yOiOj
(3)

2.2.3. Catalytic Evaluation

Catalytic ammonia synthesis tests were conducted within an Avantium system (Flowrence XD, Netherlands) using a quartz tube reactor with an I.D. of 2 mm and an O.D. of 3 mm, equipped with a porosity 3 filter. Four reactors were simultaneously employed for screening purposes, with each reactor loaded with 25 mg of catalyst, which was diluted with 600 μL of silicon carbide (SiC) grit (Lot: 10226827, Alfa Aesar). One reactor was kept as a blank control, containing only SiC. Before the activity measurement, the catalysts were treated in an in situ reduction/activation process. This was achieved under a gas mixture consisting of hydrogen (99.999%) and nitrogen (99.999%) at a 3:1 ratio, with a total flow rate of 15 mL min–1 at 650 °C (ramping rate: 5 °C min–1) for 2 h. Catalytic performance assessments were conducted at 10 bar, 400 °C. For these experiments, precise gas flow rates of 11.25 for hydrogen and 3.75 mL min–1 for nitrogen (corresponding to a weight hourly space velocity of 36,000 mL gcat–1 h–1) were maintained using mass flow controllers. An online gas chromatogram (8890 GC system, Agilent) was employed to analyze the reaction products under isothermal conditions. Helium (99.999%) at a flow rate of 1.25 mL min–1 was used as an internal standard for quantitative analysis. All experiments were replicated three times, and the average values were utilized for generating graphical representations and data analysis. To eliminate the scenario of spurious NH3 presence (e.g., due to possible impurities in the lines), control experiments were performed following the procedure: the same experimental protocol (reduction at 75% H2 −25% N2 at 650 °C for 2 h) followed by the reaction conducted in the absence of N2 in the feed (control experiment), where gas flow of H2 15 mL/min, and He 1.25 mL/min, at 400 °C, and pressure of 10 bar was used.

3. Results and Discussion

Click to copy section linkSection link copied!

3.1. Computational Studies

3.1.1. Structural Stability

Ceria exists in fluorite cubic crystal structure with space group no. 225 as shown in Figure 1a. The formation energy of ceria is −4.04 eV/atom (shown in Table 1). To understand the formation of oxygen vacancy in high-entropy systems calculations for Ce-based binary systems, CeGdO4 (coded as CeGdO), CePrO4 (coded as CePrO), ternary systems, CeLaGdO6, CeLaPrO6, (coded as CeLaGdO and CeLaPrO) and quinary high-entropy oxide CeLaPrSmGdO10 (coded as CeLaPrSmGdO) were performed. The crystal structure of the binary system is derived from the ordered arrangement of rare earth atoms in the crystal structure of ceria. For ternary and quinary high entropy systems, crystal structure is derived by SQS (22) using the ATAT package. (23) The bulk crystal structures for pure CeO2, binary CeGdO, ternary CeLaGdO, and quinary CeLaPrSmGdO are shown in Figure 1. The formation energy for binary, ternary, and quinary bulk phases is mentioned in Table 1. It is to be noted that the formation energy is negative even for high-entropy systems, suggesting that although configurational entropy contributes to the stability of these systems, it is not the governing criterion for stability, as these materials are stable even at T = 0 K.
Table 1. Formation Energy in eV/fu for CeO2 and CeO2-Based High Entropy Systems
 ΔHf (eV/atom)
CeO2–4.04
binary (CeGdO)–4.83
binary (CePrO)–3.76
ternary (CeLaGdO)–4.26
ternary (CeLaPrO)–3.54
quinary (CeLaPrSmGdO)–4.13

Figure 1

Figure 1. Unit cells of (a) pure CeO2, (b) binary CeGdO, (c) ternary CeLaGdO, and (d) quinary CeLaPrSmGdO.

3.1.2. Evolution of Oxygen Vacant Sites

To understand the evolution of oxygen vacancy formation in ceria-based high entropy systems, the oxygen vacancy formation energy (ΔEf) along (111) surface for each of the cases, pure CeO2, binary, ternary, and quinary systems was computed. As shown in Figure 2a the formation energy of oxygen vacancy defect using eq 1 in bulk, sublayer, and top-layer along (111) surface of various systems was computed. For pure CeO2 the ΔEf in bulk is 4.71 eV, on the top-layer is 2.93 eV and the sublayer is 3.21 eV, which is in agreement with previous studies. (38,39) Similarly, ΔEf is computed for different systems (shown in Figure 1) along (111) surface, values shown in Table 2 and graphically shown in Figure 2b.
Table 2. Formation Energy (ΔHf) in Top-Layer, Sub-Layer, and Bulk for Various Systems
  ΔEf (eV)
CeO2top-layer2.93
 sublayer3.21
 bulk4.71
binary (CeGdO)top-layer–0.28
 sublayer0.18
 bulk4.15
binary (CePrO)top-layer–0.24
 sublayer–0.13
 bulk2.17
ternary (CeLaGdO)top-layer–0.20
 sublayer–0.09
 bulk0.07
ternary (CeLaPrO)top-layer–0.62
 sublayer–0.50
 bulk0.29
quinarytop-layer–0.43
(CeLaPrSmGdO)sublayer0.01
 bulk0.21

Figure 2

Figure 2. (a) Unit cell of CeO2 along (110) surface with vacuum. Bulk region, sublayer, and top-layers are shown. (b) Evolution of oxygen vacancy formation energy in pure CeO2, binary systems (CeGdO, CePrO), ternary systems (CeLaGdO, CeLaPrO), and quinary (CeLaPrSmGdO) high entropy system.

Several possible sites of oxygen vacancy in bulk and on (111) surface are considered and the one with the lowest energy is considered for further analysis. It is observed that as the number of heteroatoms increases, moving from pure ceria to mixed cation systems, ΔEf reduces and eventually reaches values below zero, suggesting that in high-entropy systems (HEO) oxygen vacancy (Ov) formation is energetically more favorable as compared to pure CeO2. This is due to the ease of transition of cation from +4 oxidation state to +3 oxidation state. Pr, Sm, and Gd exhibit various oxidation states, including +3 and +4. + 3 oxidation state of La, Pr, and Gd is more favorable as they exhibit fully filled [Xe], [Xe] 6s2, and [Xe] 4f7 electron configuration. Therefore, replacing Ce with these rare-earth elements tends to ease the formation of oxygen vacancy both in bulk and on the surface. It is also noticeable that ΔEf is highest in bulk, and gradually reduces in sublayer, and is lowest at the surface. This trend is observed in all the systems. This can be explained by the fact that the oxygen atoms on the (111) surface of these ceria-based systems contain dangling bonds, such that the surface is electron-poor. Removal of an oxygen atom on the surface provides extra charge from the cations which are near to the vacancy site, to the already electron-poor surface, thereby favoring the formation of oxygen vacancy.

3.2. Experimental Studies: Monitoring of the HEO vs Binary Oxides Differences

Further understanding of the fundamental differences of the HEO, compared to (1) simpler structures of ceria-based oxides (e.g., binary supports, such as CeLaO, CePrO) and (2) nonreducible ones (Silica) was herein attempted using a toolkit of analytical techniques. For this reason, the catalysts with Ru/CeLaO, Ru/CePrO, Ru/SiO2, and Ru/HEO compositions were investigated; more details are provided in what follows.

3.2.1. Overview of Structural/Textural and Morphological Features

Figure 3A presents the XRD patterns of the Ru-based catalysts on different supports of this study following their reduction at 650 °C for 2 h under an H2 atmosphere; the supports were CePrO, CeLaO, HEO, and silica. All the ceria-based supports present the characteristic reflections of the fluorite ceria lattice; namely 33.2°, 47.6°, 56.5°, 59.1°, 69.4°, 76.6°, and 79.2° are assigned to the diffractions of (111), (200), (220), (311), (222), (400), (311), and (420) crystallographic planes of CeO2. (40) The catalyst where silica was used as support presents a rather amorphous pattern, whereas the Ru/HEO catalyst presents no reflection due to a dopant relevant phase (e.g., La2O3, Sm2O3, Gd2O3, Pr2O3, or combination/solid solution of them). On the other hand, despite the relatively high Ru loading (6 wt %), there is no obvious reflection peak corresponding to Ru/RuO2 phases, demonstrating the high dispersion of Ru and/or RuO2 phases in all the supports herein used (regardless of their nature: reducible or nonreducible). N2 adsorption at 77 K (Figure S1) over all the ceria-based supported catalysts showed similar BET surface areas (Table S1) of the catalysts Ru/CeLaO (18 m2/g), Ru/CePrO (6 m2/g), and Ru/HEO (18 m2/g), whereas the pore volume was found to be in the 0.032–0.048 cm3/g range for the ceria-based catalysts and 0.936 cm3/g for the silica-based catalyst. The Ru/SiO2 catalyst had a surface area of 102 m2/g. All the catalysts presented similar morphological features (rather aggregates formation based on the SEM studies with the Ru/SiO2 catalyst presenting a bit more spongy morphology (Figure S2). The latter is in agreement with the high surface area of this catalyst. Elemental EDS mapping measurements, by SEM, were carried out to confirm the presence of elements, and the results are presented in Table S2. Based on the values reported, the EDX-derived values for the Ce/La and Ce/Pr ratios are close to the nominal one (4/1); the same applies to the Ce/La/Sm/Gd/Pr ratio which reflects the equimolar composition of the HEO (design criterion is fulfilled).

Figure 3

Figure 3. (A) XRD diffractograms; (B–C) EPR spectra collected at 100 K; (D) Raman spectra collected over the Ru/CePrO catalyst; (E) Raman spectra collected over the Ru/HEO catalyst; (F) H2-temperature-programmed desorption (H2-TPD) profiles obtained over all the Ru catalysts of the present study.

3.2.2. Probing the Oxygen Vacant Sites and the Metal–Support Interactions

Figures 3B,C and S3 present the EPR spectra of the Ru/CeLaO (reducible oxide, reference), Ru/HEO, and Ru/SiO2 (nonreducible, reference) obtained at 100 K and room temperature, respectively, following their reduction at 650 °C for 2h in H2 atmosphere. EPR is used in this study as a diagnostic tool for the Ce3+ species; given their paramagnetic nature, Ce3+ ions can be detectable using this technique. From the catalysis perspective, it is of high importance to be able to probe the presence of oxygen vacancies (Ov); the formation of the latter is associated with the Ce3+ presence. It is worthwhile to discuss in brief the mechanisms of Ov formation in these materials; according to the first mechanistic scenario, as the lattice oxygen leaves its original position, the two electrons in the vacant site localize on two Ce4+ ions and they reduce them forming Ce3+–VO–Ce3+ pairs. (41) Usually, conventional EPR at RT (298 K) cannot detect these species due to the strong spin–spin interaction. An alternative mechanism of oxygen vacancy (Ov) formation is when one electron, originating from lattice oxygen, localizes on a Ce4+ leading to the formation of Ce4+–VO–Ce3+ species; in this scenario, the other electron is trapped in an adjacent to the vacancy position, leading to the formation of F+ center; (41) the latter can be probed by RT (298 K) EPR. In the open literature, there is a lot of discussion on the origin of peak with gav ∼ 1.97 which is usually assigned to Ce3+. Based on the literature, the symmetry around the Ce3+ ions dictates the g parameter value. As Ce4+ to Ce3+ reduction takes place, it leads to the formation of oxygen vacancies accompanied by lattice distortion. In the case of CaF2 (isomorphic) for Ce3+ ions in a coordination environment of low symmetry, g ∼ 3.67 has been reported, whereas lines with g values of 2.4 for the trigonal site of Ce3+ ions have also been reported. (41) As can be seen in Figure 3C, EPR signals at different g-values confirm the various paramagnetic centers; of high significance is the signal at g ∼2.000 which is assigned to superoxide anions (•O2–) attached to Ce4+ ions; the intensity of this pulse is correlated with the concentration of oxygen vacant sites. (42) This signal is of high intensity in the case of the Ru/HEO catalyst.
Additional structural information was acquired utilizing Raman spectroscopy over the Ru-based catalysts of this study following their reduction at 650 °C for 2h in H2 atmosphere. In particular, Raman spectroscopic results (Figure 3D,E) obtained over the Ru/CePrO and Ru/HEO catalysts give insight into the effect of dopants on the multi-elemental HEO structure and defective sites formation as well as how the two different supports interact with the metal sites (Ru) through strong metal support interactions (SMSI). A comparative Raman plot is given in Figure S4 to assist with the peaks’ assignment. It is known that the Raman spectra for typical ceria-based materials have a dominant peak, the F2g peak, usually found at about 464 cm–1 and assigned to the symmetric stretch mode of the Ce–O8 crystal unit, characteristics of the fluorite lattice structure. (43) A close look at the Raman spectra of the catalysts (Figure S4) reveals significant changes in them indicating the remarkable impact of the dopants in the ceria microstructure owing to higher defectiveness and a rise in topographical/local disorder. Particularly, the F2g peak appears quite sharp in the case of Ru/CePrO catalyst and rather suppressed in the case of Ru/HEO catalyst (see Figures 3E and S4); this is associated with the high degree of oxygen sublattice distortion in the latter case (HEO) and the high sensitivity of Raman spectroscopy to capture the oxygen sublattice disturbances. The herein Raman spectra over the HEO agree quite well with the literature reported by Sarkar et al. (44) Apart from the suppressed intensity of the F2g peak, changes in the defect-induced bands were found in the 500 to 700 cm–1 range, which was being misinterpreted in the open literature as oxygen vacancies associated with the presence of reduced Ce3+ (45−47) species or to oxygen vacancies (Ov) involving movement of an oxygen atom into an octahedral interstitial position to obtain vacancy. (48,49) The characteristic features have rightly been considered intrinsic properties of a pure ceria structure; however, intensity and broadness of the defects band are traced to extrinsic defects, attributed to the addition of dopant (Gd, Sm, La, and Pr for HEO; Pr for CePrO). (50) The abundant defects as those identified in the Raman spectra are linked to the strong interaction between ceria and doped elements, thereby obtaining a microstructural change. The introduction of dopant into the ceria microstructure causes a steady increase in the amount of intrinsic and extrinsic defects. In a more quantitative manner, this is reflected in the value of the IOv/IF2g ratio, the latter being a descriptor of the Ov abundance. In particular, the IOv/IF2g ratio equals 0.68 and 3.2 in the cases of Ru/CePrO and Ru/HEO catalysts, respectively, demonstrating the abundance of Ov in the HEO lattice (bulk) and in the medium range that is probed by Raman scattering effects. (51) Additionally, the Raman band at 1100 cm–1 corresponding to the Ru–O–Ce bond environment demonstrates strong metal–support interactions (SMSI); this band is sharp (higher signal-to-noise ratio) in the case of quinary HEO-based catalyst compared to the binary CePrO-supported one. The presence of Ru–O–Ce species is a critical factor for the Ru dispersion, as this has been assessed through H2 chemisorption studies (Figure 3F), and catalytic activity as it will be discussed later. (52,53)

3.2.3. Surface Composition and Coordination Environment

To delve deeper into the chemical states of Ru species in the as-prepared catalysts, X-ray photoelectron spectroscopy (XPS) measurements were conducted (see Figure 4A). Of particular interest in the XPS studies was to shed light on the effectiveness of the in situ Ru reduction as well as on the Ru-support interaction while Ru growth and reduction simultaneously take place during the synthesis. The peaks observed at 461.9 and 484.1 eV are attributed to metallic Ru, while those at 463.7 and 485.9 eV are associated with RuO2, as reported by Wang et al. (54) A shift to lower binding energy in the Ru 3p core level spectra, particularly when compared to the Ru 3p of the reference Ru/SiO2 catalyst suggests changes in the electronic structure or chemical environment of Ru in the ceria-supported catalysts. The surface analysis reveals an increase in Ru/Ce ratio (Ru surface enrichment) as we transition from the binary system (i.e., Ru/CePrO or Ru/CeLaO) to the high-entropy system (Ru/HEO), as shown in Table 3. This observation suggests a different interaction among the three under study supports (CePrO, CeLaO, HEO) and the Ru species, promoting a high dispersion of Ru on the surface (in the case of HEO); these results are consistent with H2-TPD (chemisorption) results and the existing literature. (55) Possible parameters contributing to the support-metal interaction are surface termination (dangling bonds, functional groups) and surface energy (J/m2). The Ru 3p spectra of Ru/CePrO, Ru/CeLaO, and Ru/HEO catalysts were fitted by CasaXPS software using Lorentzian Asymmetric line shape with Voigt function, and results are shown in Figure S5. Indeed, Ru species exist as Ru0 and Ru4+ (in RuO2), with the latter species being more dominant on the surface of the Ru/CeLaO supported catalysts (Ru4+/Ru0=15.6), as can be seen in Table 3. The ratios of Ru4+/Ru0 in Ru/CePrO, Ru/CeLaO, and Ru/HEO were found to be 5.5, 15.6, and 6.2, respectively. The increased presence of Ru4+ species in the Ru/CeLaO system, as indicated by the higher Ru4+/Ru0 ratio, suggests that the specific combination of dopants in the CeLaO system has a pronounced effect on the reducibility of Ru due to the strong metal–support interactions (SMSI).
Table 3. XPS Data of the Prepared Catalysts
catalystO (at %)Ru (at %)Ce (at %)Ru/CeRu4+/Ru0a
Ru/SiO250.530.37   
Ru/CePrO13.0124.319.822.55.5
Ru/CeLaO22.707.977.781.015.6
Ru/HEO26.583.480.93.96.2
a

Values were estimated by XPS peak fitting using CasaXPS software.

Figure 4

Figure 4. (A) Ru 3p, (B) Ce 3d, and (C) O 1s XPS core-level spectra of the prepared catalysts.

The Ce 3d spectra depicted in Figure 4B are intricate, featuring multiple peaks that are labeled based on previous XPS studies on cerium oxide. (55,56) The prominent peaks appearing at 916.2, 907, 900.3, 897.8, 888.7, and 882 eV are ascribed to Ce4+ species, whereas the two peaks observed at 902.2 and 884.6 eV are contributions of Ce3+ species. With respect to the binary systems, the Ru/HEO catalyst exhibits a shift of binding energy in higher values indicating a change in the Ce environment upon doping. Overall, a remarkable decrease in surface Ce species is observed (see Table 3), as well as intensity increase in Ce3+-related peaks can be visualized based on its spectrum in Figure 4B, likely attributed to the formation of extrinsic defects like oxygen vacancies due to doping with multiple elements.
In the O 1s spectra (Figures 4C and S5), the peak at approximately ∼528.8 eV is attributed to lattice oxygen, while the peaks around ∼530 eV, 531, and 533 eV are linked to oxygen vacancies, surface hydroxyl species, and carbonates, respectively. (57) A significant change in the O environment is noticed across the catalysts. Moving from binary systems (e.g., CePrO and CeLaO) to the HEO system reveals increased peak intensities. This indicates that varying the chemical compositions has an impact on the type and concentration of surface oxygen species due to the formation of distinct metal oxides or mixed-metal oxides. Ru/HEO presents the highest content of surface oxygen species, likely attributed to the optimal formation of Ru–O–Ce bonds. Deconvolution of the O 1s peak (Figure S6 and Table S3), although, revealed that for the Ru/CePrO, Ru/CeLaO, and Ru/HEO catalysts the lattice oxygen (OL) abundance on the surface is 20, 28, and 42%, respectively. This result leads to the conclusion that the Ru/HEO catalyst has the least Ov surface concentration. Caution should be exercised in the interpretation of this data due to the fact that these values are only a rough estimation and will probably not give the correct stoichiometry, which presumably is (Ce, X, Y, Z)O2 in the case of HEO. Furthermore, the higher abundance of Olattice (associated with lower abundance of Ov) in the case of HEO catalyst can be understood by taking into consideration the following: (1) the XPS studies were performed on the as-prepared catalyst; (2) the XPS experiment was of ex-situ nature; (3) the fact that Ov creation can happen under the vacuum-catalyst interaction (ultrahigh vacuum in the XPS experiment) and (4) XPS experiment captures only the surface, so Ov at the different locations (subsurface/bulk) escape the detection depth of XPS. Additionally, the dynamics of Ov is a parameter to not be ignored. As pointed out by Younis et al. (58) at high levels of doping (>20%) the vacancies become immobile due to their clustering and this can lead to deterioration of catalytic activity, as it will be discussed later.

3.2.4. Coordination Environment through Synchrotron EXAFS Studies

Long and medium-range features, as those can be derived using XRD (long-range) and Raman (medium), can be hardly used to describe structural differences for materials in the nm-range. Local distortion with no periodicity can dictate the local structure and the ultimate functionality of a material. (59) Synchrotron EXAFS was used to study the Ce–O correlations in a changing environment from the binary Ce–Pr oxide to the HEO. (60)
The linear combination fitting (Figure 5A,B) of the XANES data collected at the L3 edge of Ce reveals a mixture of Ce3+ and Ce4+ oxidation states. In particular, the dominating oxidation state in the Ru/CePrO catalyst is Ce3+ at an abundance of ∼96%; in the case of the Ru/HEO catalyst, the Ce3+/Ce4+ ratio is about 1.2. The above results are important, though should be interpreted given the short-range sensitivity of the technique (different length scale with the focus being primarily around the atoms of interest).

Figure 5

Figure 5. (A, B) Linear combination fitting of the XANES spectra collected at the LIII edge of Ce in the HEO sample (left) and CePrO sample (right) to determine all the oxidation states of Ce in the samples; (C) Fourier transforms with their respective real part of the k3 weighted EXAFS spectra collected at the Ce LIII edge in CLPSG-HEO and CePr samples.

Analyzing a short EXAFS signal can give good information on the first coordination shell around Ce (nearest neighbor, NN). Derived structural parameters from the EXAFS fitting show a coordination number of about 7 oxygen atoms with an interatomic distance that is close to that of CeO2 with a slight contraction in CePr to 2.30 Å and even shorter in the case of the HEO sample. In particular, the Ce–O distance becomes smaller as we move from ceria (2.34 Å) to CePrO (2.30 Å) to HEO (2.27 Å) (Table 4). The short distance coincides also with a reduction of the coordination number from 8 oxygen atoms for CeO2 to an average number of 7.4 and 7.1 atoms in CePr and HEO samples, respectively. These results may suggest a defect in the local structure at short-medium (could be defects or distortion). It should be mentioned here that, it was not possible to collect long EXAFS data (signal) at the L3 edge of Ce (due to the presence of the L2 edge at around 400 eV), and thus the coordination number is largely affected by the noise level which is reflected in the higher standard deviation (2.4 for N and 0.036 for the R, Table 4). This can be considered as a limitation of the above data, making EXAFS results only complementary to the above XANES ones. Moreover, based on the above HEO seems to be evolving as a triclinic CeO2 which has a shorter Ce–O distance (2.27) than the CeO2 cubic (2.34).
Table 4. Structural Parameters Derived by EXAFS Fitting of the First Shell around the Ce Atom in CLPSG-HEO and CePr Samples
samplebondN (atom)R (Å)σ2 (Å2)ΔE(eV)
HEOCe–O7.1 ± 2.42.27 ± 0.0360.01460 ± 0.0043.99 ± 1.6
CePrOCe–O7.4 ± 2.02.30 ± 0.0100.01542 ± 0.0014.05 ± 0.5
CeO2 (cubic)Ce–O8a2.34248a  
a

Theoretical value.

3.2.5. Ru Particle Shape and Size on the Different Supports (Electron Microscopy and Chemisorption Studies)

In the present study, the Ru dispersion and particle size were evaluated using H2 chemisorption studies, where the H atoms dissociative adsorb onto the Ru metal with a H:Ru stoichiometry of 1:1. Additionally, HRTEM analysis was employed to verify the chemisorption studies. Figure 3F displays the H2-TPD profiles of the Ru-based catalysts of this study. According to the open literature, the peak at low temperature (75 °C, 135 °C) is assigned to the desorption of hydrogen from various active sites (metallic Ru of different sizes), whereas the peak with Tmax > 380 °C corresponds to the spillover hydrogen from the support. The corresponding Ru particle size was found to be in the 4–11 nm range demonstrating different extents of Ru dispersion as shown in Table S4. This Ru particle size variation, as probed using chemisorption, is expected to have a profound impact on the NH3 synthesis catalytic activity due to the well-known structural and size sensitivity of the reaction at hand and particularly of the N2 activation step. (61)
In a complementary fashion to the H2 chemisorption studies, HRTEM studies were performed and the results are presented in Figures 6, 7, 8, 9, and 10 along with their STEM-HAADF Red Green Blue (RGB) and Fast Fourier Transform (FFT) analysis which allows us to comment on the distribution of the elements in the oxide matrix and the exposed facets in the particles, respectively. In particular, in Figures 6 and 7, and Figures 8, 9, and 10 the HRTEM and STEM-HAADF RGB analysis of the Ru/CePrO and Ru/HEO catalysts, respectively, are presented, following their reduction at 650 °C for 2 h. Comparing the HRTEM images in Figure 6 with Figures 8 and 9 it can be concluded that Ru particles grow in a more faceted manner (sharp planes) over the CePrO compared to HEO support. In the Figure 6, lattice fringes with interplanar spacing of 0.196 nm, 0.227 nm can be assigned to the Ru(101) and Ru(100) planes, respectively.

Figure 6

Figure 6. (A) HRTEM images along with Fast Fourier Transform (FFT) pattern; (B, C) facets analysis of the Ru/CePrO catalyst.

Figure 7

Figure 7. (A–E) STEM-HAADF RGB analysis and (F) selected area electron diffraction (SAED) over the Ru/CePrO catalyst.

Figure 8

Figure 8. (A–C) HRTEM images of the Ru/HEO catalyst with emphasis on Ru particles at different areas (B) and (C).

Figure 9

Figure 9. (A) HRTEM images of the Ru/HEO catalyst along with Fast Fourier Transform (FFT) pattern and (B–D) facets analysis.

Figure 10

Figure 10. (A–H) STEM-HAADF RGB analysis and (I) selected area electron diffraction (SAED) over the Ru/HEO catalyst.

Ammonia synthesis reaction and its rate-determining step, that of N2 activation, exhibits high sensitivity in particle shape and faceting. (62) The facets exposed in the cases of Ru/CePrO and Ru/HEO were confirmed through FFT analysis (Figures 7 and 9 and Tables S5, S6). Regarding the Ru particle size, histogram analysis (Figure S7), based on the HRTEM studies, also showed that Ru particle size is ∼6 and ∼12 nm in the cases of HEO and CePrO-supported catalysts, respectively. These values of particle size are quite in agreement with the H2 chemisorption studies above; some overestimation of dispersion can be observed in the chemisorption studies due to the well-known phenomenon of atomic hydrogen spillover in the ceria-based supports. In brief, it is reported by Rarog-Pilecka et al. (63) that the ammonia synthesis rate increased with the particle size increasing from 0.7 to 4 nm, whereas smaller sizes of Ru crystallites (smaller than 0.7 nm) can be totally inactive; this is linked to the lack of B5 sites. Kim et al. reported that double-stepped Ru(109) sites are more active than the stepped Ru(0001) surface. (64−70)

3.2.6. Catalytic Performance toward NH3 Production

All the Ru-based catalysts of this study were tested following their reduction at 650 °C for 2h for ammonia production at different reaction pressures (Figure 11A), whereas the best-performing one was evaluated following different activation conditions (Figure 11B). In particular, reaction pressures in the 10–50 bar were used; additionally, the catalysts were subjected to different activation temperatures so to induce different populations of Ov sites; namely, reduction at 650 °C (1) and reduction at 800 °C (2). The Ru/HEO catalyst following reduction at 650 °C (catalyst 1, Figure 11A) led to the production of 1236 μmols NH3/g-h at a pressure of 10 bar. Increasing the reaction pressure results in a drop in catalytic activity for the Ru/HEO catalyst, whereas the rest of the catalysts presented activity close to zero. Following its activation/reduction to 800 °C for 2 h and keeping the reaction pressure at 10 bar, the NH3 production drops to 954 μmols NH3/g-h (Figure 11B). It is expected that two major parameters are playing a key role in the measured catalytic activity, as presented in Figure 11 toward ammonia production; Ru particle size as this is controlled by the sintering extent (650 vs 800 °C) and active sites population as this dictated by the Ru particle shape; additionally, it is expected that oxygen vacant sites from the HEO are also a critical activity descriptor. It is well documented that N2 dissociation (bond energy 945 kJ/mol) is the rate-limiting step in the NH3 production. (68,61) Under reduction conditions it is expected that electron donation is enhanced from Ru to the π* antibonding orbital of N2 facilitating the N2 activation.

Figure 11

Figure 11. Catalytic production of NH3 (A) over Ru-based catalysts of this study in the 10–40 bar pressure range and 400 °C; (B) over the Ru/HEO catalyst at 10 bar, 400 °C following two different activation conditions (650 vs 800 °C).

3.2.7. Intertwined Activity Descriptors

In particular, it has been demonstrated by Jacobsen et al. (61) that the Ru particle size in the 1.8–2.5 nm range contains the maximum number of B5 sites; the latter are the ones favoring the N2 dissociative adsorption. It is believed that the results of Figure 11 are the outcome of two competitive trends; the B5 sites drop in population due to Ru sintering and the Ov increase in population due to the reduction conditions (Ov-650 °C > Ov-800 °C). The trend observed in NH3 production is in agreement with the findings of the 18O2 transient isotopic isothermal experiments and the population of the surface Ov (discussed later). For study completeness, and to ensure that there is no NH3/N2 impurity that can affect the obtained measurements of our study (overestimate of ammonia production), the above experiments were also run under the same conditions in the absence of N2 in the feed; no NH3 production was found, thus confirming the lack of ambiguity in our measurements as per the open literature. (69,70)

3.2.8. Ru Metal Particles’ Role

In an effort to untangle the role of Ru particle size on the N2 dissociation, it is worth revisiting the recently published report by Yanliang Zhou, (71) where it is clearly stated that Ru size reduction, in the case of Ru/BaCeO3 catalysts, enhances the formation of Ce3+ species and oxygen vacancies (Ov) sites; the latter facilitates the donation of electrons to the Ru sites and thus enhances the N2 dissociation step. Additionally, the structure sensitivity of the ammonia synthesis reaction has been highlighted in the study of Peng et al. (72) at mild conditions; it is manifested by significant changes in catalyst activity at only small structural alterations. In particular, it was reported that the population of corner sites (low coordination sites) increases with dropping Ru particle size with a simultaneous drop of the terrace sites population; additionally, due to the geometric changes (shape and size), the Ru electronic structure is tailored with its size tuning. This causes a reduction to the catalyst work function, ϕcat, which subsequently facilitates the electron donation from the d-orbitals of Ru to the N2, enhancing N2 activation and N–H bond formation. Isotopic labeling experiments coupled with DRIFTS can be used to trace the size of the pool of the key N-containing intermediates and the mechanism followed (associative vs dissociative); this is important, particularly due to the fact that Ru size can be the crucial factor determining the mechanism too; namely, Ru-based catalysts with particle size >2 nm favor a dissociative pathway, (73) whereas associative pathway is followed in the case of almost atomically dispersed Ru catalysts. (74,75) Moreover, smaller Ru size (higher dispersion) can facilitate the H-spillover from Ru to support/interface and induce the H trapping in Ov-H entities, thus blocking the H-poisoning (parasitic) effect of Ru (76) which takes place under the reaction conditions. It has to be mentioned that the location (surface/subsurface) and dynamics (clustering) of the Ov in trapping H atoms in Ov-H entities are yet to be determined.

3.2.9. Oxygen Vacancies’ Role

It is noteworthy that Ov plays a crucial role and it is almost synonymous with the ceria-based materials’ functionality. However, caution should be exercised when a direct correlation of Ov presence and catalytic activity is attempted for reasons commented in what follows and related to Ov location and dynamics. Jennifer L. M. Rupp and her collaborators in an enlightening review (77) discuss the structural arrangements of the Ov in the ceria-doped materials (Ce3+ = 128 pm, Ce4+ = 111 pm). Authors point out that elements/dopants smaller than Gd (Gd3+ = 119 pm) favor the placement of the vacancy at the nearest neighbor (NN) site, whereas the elements that are larger than Sm (Sm3+ = 96 pm) repel the vacancy from this location. Themselves, Gd and Sm, exhibit neither attractive nor repulsive traits toward vacancies location in the NN site. An opposite trend has been found for the case of the next nearest neighbor (NNN) site, where locking the Ov at this site seems never to be preferred. Computational studies show that as the population of NN rare earth ions surrounding the vacancy increases, the defect association energy increases too. Therefore, the local configurations of dopants and oxygen vacancies in doped ceria are of particular importance, yet some details are still not completely understood. Particularly for the N2 reduction step, there are many studies trying to unresolve the Ov role in the reaction. In the work by Yanliang Zhou (71) authors investigated the effect of different contents of Ov (equivalent to different contents of Ce3+ entities) through Bader charge analysis. In the structural models (zero Ov, 1 Ov, and 2 Ov) adopted, the Ov content was 0, 3.3, and 6.7%, respectively, giving rise to 2.3–5% Ov content in the Ru/BaCeO2 catalyst. It was found that the charge donation from Ce3+ and Ov to the Ru increases as the Ov content increases. Oxygen vacancies dictate the charge accumulation to Ru particles, while they can be the drive for mechanism pathway change. (78) Though, another characteristic of Ov that needs to be discussed herein is their clustering at a certain level of doping. Younis et al. report the drop in the catalytic activity of Gd-doped ceria when the dopant level passes 15% where the Ov vacancies become immobile due to their clustering, stressing the importance of vacancies dynamics. (58)

3.3. Insights into the Oxygen Mobility and Vacancy Formation

To get a thorough understanding of the fine differences in terms of oxygen mobility and vacancy formation of the HEO vs CePrO binary oxide, a 18O2 transient isotopic isothermal exchange (TIIE) experiment was performed. The transient rates of 16O2 consumption (μmol g–1 s–1) obtained during TIO at 650 °C followed by catalyst pretreatment (Exp 1–3) over the HEO solid, are shown in Figure 12A. It can be clearly seen that the rate of oxygen consumed (and its maximum) is lower in the case where no reduction (Exp 1) was performed during the pretreatment of the solid, compared to the cases where H2 reduction was performed. It is evident that by applying H2 reduction after calcination, oxygen vacancies were formed, the amount of which depends on the temperature of reduction. At the initial stage (before maximum), Exp 2 and Exp 3 exhibit similar rates, indicating similar amounts of surface oxygen vacancies. Whereas, with time under 16O2/Kr/Ar/He gas stream, the amount of oxygen vacancies in the bulk is higher in the case where the catalyst reduced at 650 °C, instead of 800 °C. It should be noted that the higher the O2 consumption amount, the more oxygen vacancies exist. Figure 12B presents the total amount of O2 consumed (NV, μmol g–1) as a function of pretreatment conditions applied over the two solids. More precisely, independent of the catalyst pretreatment protocol followed, the HEO exhibits more oxygen vacant sites, compared to the CePrO, with an optimum condition being that of catalyst reduction with 5% H2/He at 650 °C (Exp 2).

Figure 12

Figure 12. (A) Transient rates of 16O2 (μmol g–1 s–1) consumed during transient isothermal oxidation (TIO) at 650 °C, following the different pretreatment conditions, over HEO solid; and (B) amounts of O2 consumed (μmol g–1) during TIO, over both solids.

Figure 13A displays the transient concentration (mol %) response curves of Kr, 16O2, 16O18O, and 18O2 obtained during the 18O2-TIIE switch from 16O2/Kr/Ar/He to 18O2/Ar/He gas mixture at 650 °C, over HEO pretreated via Exp 1. After the switch to the 18O2/Ar/He, the 16O2(g) and 18O2(g) concentrations decrease and increase, respectively, while the formation of 16O18O(g) starts immediately and passes through a maximum at about 450 s, following by a decreasing tail controlled mostly from the bulk 16O diffusion toward the surface. The large difference between 16O2 decay and that of inert Kr, which lasts for ∼15 s, can be attributed to the surface 16O, 18O, and 16O18O exchange on the catalytic surface, considering that the amount of 16O exchanged reached the ML = 1 in the first 20 s of the transient. For t > 20 s, subsurface and bulk 16O diffusion and exchange toward 16O18O formation is dominant. The total amount of 16O18O exchanged (N16O), estimated via eq 2, was found ∼1.2 times lower in the case of HEO compared to CePrO, while increases by increasing the H2 reduction temperature (from 650 to 800 °C) during the pretreatment process (9.75, 10.21, and 11.99 mmol g–1 for HEO, respectively, compared to 11.92, 12.02, and 12.48 mmol g–1 for CePrO).

Figure 13

Figure 13. (A) Transient concentration (mol %) response curves of Kr, 16O2, 16O18O, and 18O2 gaseous species recorded during the 18O2-TIIE experiment at 650 °C, over the calcined HEO; (B) αg(18) descriptor as a function of time and pretreatment procedure, estimated during 18O2-TIIE over HEO; and (C) αg(18) descriptor obtained over the both reduced at 650 °C solids.

The effect of H2 reduction temperature on the αg(18)(t) descriptor function, over the HEO is illustrated in Figure 13B. It can be seen that different shapes and positions are obtained in the whole transient by varying the catalyst pretreatment (similar results were obtained over CePrO, not shown). Getting more specific, at the initial period after the 18O2/Ar/He gas switch, where the 16O/18O exchange on the surface takes place, a steeper response is obvious in the absence of H2 pretreatment (Exp 1), indicating a higher Deff value, (37) opposed to the Exp 2 and 3, where H2 reduction was applied after catalyst calcination. The latter behavior could be partly attributed to catalyst sintering during reduction, limiting the surface 16O/18O exchange. However, for prolonged times under the 18O2/Ar/He gas mixture (t > 450 s), it is clear that when the catalyst is reduced at 650 °C, enhances the oxygen diffusion in the bulk, compared to Exp 3 and 1. Based on the above-offered discussion, and comparing the HEO and CePrO pretreated with 5% H2 at 800 °C (see Figure 13C), it is rather clear that the former catalyst promotes to a higher extent the surface 16O/18O exchange (for t < 450 s, steeper and lower αg(18)), opposed to the bulk oxygen diffusion (t > 450 s), which is higher for the latter Ce20Pr catalytic system.

3.3.1. Insights into the Adsorption of N2/N Species on Ru/HEO Catalysts

The understanding of the interaction between the Ru/HEO catalytic surface and Nx (x = 1 or 2) species is expected to avail valuable insights into ammonia synthesis reaction and the experimental results presented above (Figure 13). Therefore, the ab initio studies were performed to examine the catalytic activity of the Ru nanoparticles and HEO support in the adsorption of N2 molecules in the presence and absence of hydrogen. Considering the first step for ammonia synthesis is the activation of the N2 molecule, the adsorption of the N2 molecule on the reduced HEO surface (Ru4/HEO) is considered by placing the N2 molecule at different sites on the surface consisting of Ru4 cluster. The configuration with the minimum energy is considered for further analysis and is shown in Figure 14a. As shown in the figure the N2 molecule prefers to bind to the surface over the Ru4 cluster. This suggests that the presence of the Ru4 cluster facilitates the adsorption of N2 molecules on the HEO surface, thereby initiating the first step in ammonia formation. The adsorption energy of the N2 molecule is calculated to be −0.35 eV, with a Ru–N bond length of 2.02 Å. The N–N bond length decreased slightly to 1.17 eV compared to the N–N bond length in the gas phase (1.19 Å). For the next step in ammonia synthesis, the adsorption of N2 molecules in the presence of H2 is analyzed. The adsorption energy of N2 further reduces to −0.68 eV in the presence of H2 compared to −0.35 eV in the absence of H2. The adsorption configuration is shown in Figure 14b. In this step, the N≡N is broken into double bonds, with top N bonded to H with a bond length of 0.89 Å and bottom N bonded to the Ru4 cluster with a bond length of 2.02 Å.

Figure 14

Figure 14. Top and side views of N2 Adsorption configurations on Ru4/reduced HEO surface: (a) in absence of hydrogen and (b) in the presence of hydrogen. Gd is shown in purple, La in dark green, Pr in yellow, Sm in pink, Ce in light green, Ru in orange, N in white, H in blue, and O in red.

It is expected that upon the N2 adsorption onto the Ru/HEO catalyst, N2 is being dissociated onto Ru sites followed by N≡N bond scission and N atom spillover onto the HEO support. Due to the above findings and including the N spillover effect the binding strength of atomic N on reduced quinary CeLaPrSmGdO (111) surface (i.e., HEO surface) is investigated. The adsorption of atomic N over a reduced HEO surface is studied by placing N over different sites over the reduced HEO surface. Our analysis shows that N over the oxygen-vacant site is most energetically favorable with an adsorption energy of −0.70 eV. The adsorption configuration is shown in Figure 15a. The adsorption of atomic N over the HEO surface is also studied in the presence of H2. The adsorption energy of N2 over the HEO surface in the presence of H2 reduced to −1.18 eV with the H atom bonded to the surface N atom with an N–H bond length of 0.89 Å.

Figure 15

Figure 15. Top and side views of N Adsorption configurations on reduced quinary CeLaPrSmGdO (111) surface: (a) in absence of hydrogen and (b) in the presence of hydrogen. Gd is shown in purple, La in dark green, Pr in yellow, Ce in light green, Sm in pink, H in blue, N in white, and O in red.

4. Conclusions

Click to copy section linkSection link copied!

Briefly, the main conclusions that can be withdrawn from the study are 2-fold. First, in terms of the understanding of the structure of the HEO compared to reference binary oxides, CePrO and CeLaO, the following remarks can be pointed out based on the multirange analysis that was performed, XRD (long-range), Raman (medium range) and EXAFS (short range): (1) according to the XRD studies, the HEO reported herein seems to adopt a cubic lattice being isostructural with the CePrO and CeLaO (reference, reducible oxides); (2) EPR and Raman demonstrated that HEO structure bears higher population of oxygen vacancies (Ov) compared to the binary CePrO under operating conditions (reduction at 650 °C/2 h); though, the abundance of Ov varies depending on the length scale probed and the surface/subsurface. In the case of Ru-supported catalysts over the HEO and CePrO supports the quinary one induces strong metal–support interactions (SMSI) and those are manifested through the formation of Ru–O–Ce interfacial species to a greater extent than in the binary oxide (Raman), and the different Ru/Ce and Ru4+/Ru0 surface intrinsic ratios (XPS); (3) HRTEM studies revealed the effect of support (HEO vs CePrO) on the different mode of growth of Ru particles shape and size (geometrical factor) and demonstrated the role of HEO in getting smaller Ru particle sizes ∼6 nm which are less sharply faceted; (4) In order to induce different degree of Ov formation in the oxide lattice, different conditions of H2 reduction, after calcination, were applied, driving the formation of Ov, the amount of which depends on the temperature of reduction (600 vs 800 °C); (5) HEO exhibits more oxygen vacant sites, compared to the reference binary CePrO, yet reducible oxide, with an optimum condition being that of catalyst reduction with 5% H2/He at 650 °C due to the extensive sintering happening at 800 °C; (6) The total amount of 16O18O exchanged (N16O), was found ∼1.2 times lower in the case of HEO compared to CePrO, while increases by increasing the H2 reduction temperature (650 °C, and 800 °C) during the pretreatment process; (7) HEO promotes in a higher extend the surface 16O/18O exchange (for t < 450 s, steeper and lower αg(18)), opposed to the bulk oxygen diffusion (t > 450 s), which is higher for the CePrO catalytic system.
Second, these Ru-based catalysts were evaluated toward their activity for ammonia production; the latter being a probe reaction targeting to unveil important features of the above catalysts and guide potential catalysts’ design. In terms of the role of the HEO in a Ru/HEO catalyst for ammonia production the following remarks can be made: (1) In the HEO, the higher amount of oxygen vacancies formed at 650 °C, and higher extend of Deff, promotes the surface 16O/18O exchange and leads to surface Ov (potential active sites for N2 and H2 activation); (2) the adsorption of N atom (being spilled over from the Ru) is more favorable at the neighboring site to oxygen vacancy (location of oxygen vacancy) as compared to the one far from the Ov; (3) The intertwined connection of Ru particle size and population/local environment of surface Ov in the HEO facilitates the efficient N2 dissociation and thus leads to higher NH3 yield (μmol/g-h); (4) The different degree of Ru particles faceting is associated with the variation of B5 sites population and thus the different performance of the catalysts for the reaction at hand.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsami.3c16521.

  • Experimental protocol for catalysts characterization that was followed in Raman spectroscopy; X-ray photoelectron spectroscopy; high-resolution transmission electron microscopy (HRTEM), electron paramagnetic resonance (EPR), and H2 chemisorption (H2-temperature-programmed desorption, TPD); textural analysis results as obtained after N2 adsorption at 77 K; SEM microphotographs obtained over the Ru catalysts of this study; EPR spectra obtained at room temperature (298 K) over 6Ru/SiO2, 6Ru/CePrO, and 6Ru/HEO catalysts; comparative raman spectra of Ru/CePrO and Ru/HEO catalysts; deconvolution of Ru 3p core-level spectra of Ru/CePrO and Ru/HEO; deconvolution of O 1s core-level spectra of Ru/CePrO, Ru/CeLaO and Ru/HEO catalysts; areas of the O 1s deconvoluted peaks; Ru dispersion and particle size as obtained from the H2 chemisorption studies; and FFT analysis of the Ru/HEO catalyst (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
    • Nirpendra Singh - Center for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab EmiratesPhysics Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab EmiratesOrcidhttps://orcid.org/0000-0001-8043-0403 Email: [email protected]
    • Kyriaki Polychronopoulou - Mechanical Engineering Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab EmiratesCenter for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab EmiratesOrcidhttps://orcid.org/0000-0002-0723-9941 Email: [email protected]
  • Authors
    • Omer Elmutasim - Mechanical Engineering Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab EmiratesCenter for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    • Aseel G. Hussien - Mechanical Engineering Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab EmiratesCenter for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    • Abhishek Sharan - Center for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab EmiratesPhysics Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    • Sara AlKhoori - Mechanical Engineering Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab EmiratesCenter for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    • Michalis A. Vasiliades - Department of Chemistry, Heterogeneous Catalysis Laboratory, University of Cyprus, 1 University Avenue, University Campus, 2109 Nicosia, CyprusOrcidhttps://orcid.org/0000-0002-1568-1158
    • Inas Magdy Abdelrahman Taha - Physics Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates
    • Seokjin Kim - Oxide & Organic Nanomaterials for Energy & Environment (ONE) Laboratory, Advanced Membranes & Porous Materials (AMPM) Center, and KAUST Catalysis Center (KCC), Physical Science & Engineering (PSE), King Abdullah University of Science and Technology (KAUST), Thuwal 23955, Saudi ArabiaOrcidhttps://orcid.org/0000-0002-6052-547X
    • Messaoud Harfouche - Synchrotron-Light for Experimental Science and Applications in the Middle East (SESAME), Allan 19252, Jordan
    • Abdul-Hamid Emwas - Core Laboratories, King Abdullah University of Science and Technology (KAUST), Thuwal 23955-6900, Saudi Arabia
    • Dalaver H. Anjum - Center for Catalysis and Separation (CeCaS), Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab EmiratesPhysics Department, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab EmiratesOrcidhttps://orcid.org/0000-0003-2336-2859
    • Angelos M. Efstathiou - Department of Chemistry, Heterogeneous Catalysis Laboratory, University of Cyprus, 1 University Avenue, University Campus, 2109 Nicosia, CyprusOrcidhttps://orcid.org/0000-0001-8393-8800
    • Cafer T. Yavuz - Oxide & Organic Nanomaterials for Energy & Environment (ONE) Laboratory, Advanced Membranes & Porous Materials (AMPM) Center, and KAUST Catalysis Center (KCC), Physical Science & Engineering (PSE), King Abdullah University of Science and Technology (KAUST), Thuwal 23955, Saudi ArabiaOrcidhttps://orcid.org/0000-0003-0580-3331
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

The authors highly acknowledge the support of Khalifa University through the RC2-2018-024 grant as well as the support from Khalifa University’s high-performance computing and research computing facilities.

References

Click to copy section linkSection link copied!

This article references 78 other publications.

  1. 1
    Mehta, P.; Barboun, P.; Herrera, F. A.; Kim, J.; Rumbach, P.; Go, D. B.; Hicks, J. C.; Scheider, W. F. Overcoming Ammonia Synthesis Scaling Relations with Plasma-enabled Catalysis. Nat. Catal. 2018, 1, 269275,  DOI: 10.1038/s41929-018-0045-1
  2. 2
    Guo, C.; Ran, J.; Vasileff, A.; Qiao, S. Z. Rational Design of Electrocatalysts and Photo(electro)catalysts for Nitrogen Reduction to Ammonia (NH3) under Ambient Conditions. Energy Environ. Sci. 2018, 11, 4556,  DOI: 10.1039/C7EE02220D
  3. 3
    Kordali, V.; Kyriacou, G.; Lambrou, Ch. Electrochemical Synthesis of Ammonia at Atmospheric Pressure and Low Temperature in a Solid Polymer Electrolyte Cell. Chem. Commun. 2000, 16731674,  DOI: 10.1039/b004885m
  4. 4
    Tao, H.; Choi, C.; Ding, L. X.; Jiang, Z.; Han, Z.; Jia, M.; Fan, Q.; Gao, Y.; Wang, H.; Robertson, A. W.; Hong, S.; Jung, Y.; Liu, S.; Sun, Z. Nitrogen Fixation by Ru Single-Atom Electrocatalytic Reduction. Chem. 2019, 5, 204214,  DOI: 10.1016/j.chempr.2018.10.007
  5. 5
    Yao, H. C.; Yao, Y. F. Ceria in Automotive Exhaust Catalysts: I. Oxygen Storage. J. Catal. 1984, 86, 254265,  DOI: 10.1016/0021-9517(84)90371-3
  6. 6
    Petallidou, K. C.; Polychronopoulou, K.; Boghosian, S.; Rodriguez, S. G.; Efstathiou, A. M. Water–Gas Shift Reaction on Pt/Ce1–xTixO2−δ: The Effect of Ce/Ti Ratio. J. Phys. Chem. C 2013, 117, 2546725477,  DOI: 10.1021/jp406059h
  7. 7
    Dou, B.; Zhao, L.; Zhang, H.; Wu, K.; Zhang, H. Renewable hydrogen production from chemical looping steam reforming of biodiesel byproduct glycerol by mesoporous oxygen carriers. Chemical Engineering Journal 2021, 416, 127612,  DOI: 10.1016/j.cej.2020.127612
  8. 8
    Kim, Y. M.; Lohsoontorn, P. K.; Bae, J. Effect of Unsintered Gadolinium-doped Ceria Buffer Layer on Performance of Metal-supported Solid Oxide Fuel Cells Using Unsintered Barium Strontium Cobalt Ferrite Cathode. J. Power Sources 2010, 195, 64206427,  DOI: 10.1016/j.jpowsour.2010.03.095
  9. 9
    Sarkar, A.; Velasco, L.; Wang, D.; Wang, Q.; Talasila, G.; Biasi, L. D.; Kubel, C.; Brezesinski, T.; Bhattacharya, S. S.; Hahn, H.; Breitung, B. High Entropy Oxides for Reversible Energy Storage. Nat. Commun. 2018, 9, 3400,  DOI: 10.1038/s41467-018-05774-5
  10. 10
    Albedwawi, S. H.; Aljaberi, A.; Haidemenopoulos, G. N.; Polychronopoulou, K. High Entropy Oxides-Exploring a Paradigm of Promising Catalysts: A Review. Materials & Design 2021, 202, 109534,  DOI: 10.1016/j.matdes.2021.109534
  11. 11
    Chae, S.; Williams, L.; Lee, J.; Heron, J. T.; Kioupakis, E. Effects of Local Compositional and Structural Disorder on Vacancy Formation in Entropy-stabilized Oxides from First-principles. npj Comput. Mater. 2022, 8, 95,  DOI: 10.1038/s41524-022-00780-0
  12. 12
    Ye, S.; Zhu, J.; Zhu, S.; Zhao, Y.; Li, M.; Huang, Z.; Wang, H.; He, J. Design strategies for perovskite-type high-entropy oxides with applications in optics. ACS Appl. Mater. Interfaces 2023, 15, 4747547486,  DOI: 10.1021/acsami.3c09447
  13. 13
    Sarkar, A.; Kruk, R.; Hahn, H. Magnetic properties of high entropy oxides. Dalton Transactions 2021, 50, 19731982,  DOI: 10.1039/D0DT04154H
  14. 14
    Sarkar, A.; Wang, Q.; Schiele, A.; Chellali, M. R.; Bhattacharya, S. S.; Wang, D.; Brezesinski, T.; Hahn, H.; Velasco, L.; Breitung, B. High-entropy oxides: Fundamental aspects and electrochemical properties. Adv. Mater. 2019, 31, e1806236  DOI: 10.1002/adma.201806236
  15. 15
    Brahlek, M.; Gazda, M.; Keppens, V.; Mazza, A. R.; McCormack, S. J.; Gryn, A. M.; Musico, B.; Page, K.; Rost, C. M.; Sinnott, S. B.; Toher, C.; Ward, T. Z.; Yamamoto, A. What is in a Name: Defining “High Entropy” Oxides. APL Mater. 2022, 10, 110902  DOI: 10.1063/5.0122727
  16. 16
    Kresse, G.; Furthmuller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 1116911186,  DOI: 10.1103/PhysRevB.54.11169
  17. 17
    Kresse, G.; Furthmuller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 1550,  DOI: 10.1016/0927-0256(96)00008-0
  18. 18
    Pedrew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
  19. 19
    Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953,  DOI: 10.1103/PhysRevB.50.17953
  20. 20
    Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-wave Method. Phys. Rev. B 1999, 59, 1758,  DOI: 10.1103/PhysRevB.59.1758
  21. 21
    Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.; Sutton, A. P. Electron-energy-loss Spectra and the Structural Stability of Nickel Oxide: An LSDA+U Study. Phys. Rev. B 1998, 57, 1505,  DOI: 10.1103/PhysRevB.57.1505
  22. 22
    Zunger, A.; Wei, S. H.; Ferreira, L. G.; Bernard, J. E. Special Quasi-random Structures. Phys. Rev. Lett. 1990, 65, 353
  23. 23
    Walle, A. V.; Tiwary, P.; Jong, M. D.; Olmsted, D. L.; Asta, M.; Dick, A.; Shin, D.; Wang, Y.; Chen, L. Q.; Liu, Z. K. Efficient Stochastic Generation of Special Quasi-random Structures. Calphad 2013, 42, 1318,  DOI: 10.1016/j.calphad.2013.06.006
  24. 24
    Ghosh, G.; Walle, A. V.; Asta, M. First-principles Calculations of the Structural and Thermodynamic Properties of bcc, fcc and hcp Solid Solutions in the Al–TM (TM = Ti, Zr and Hf) Systems: A Comparison of Cluster Expansion and Supercell Methods. Acta Mater. 2008, 56, 3202,  DOI: 10.1016/j.actamat.2008.03.006
  25. 25
    Qu, P. F.; Wang, G. C. Theoretical Insight into the Strong Size-dependence of Dry Reforming of Methane Over Ru/CeO2. J. CO2 Util. 2022, 65, 102221,  DOI: 10.1016/j.jcou.2022.102221
  26. 26
    Zafar, H.; Albedwawi, S.; Polychronopoulou, K.; Anjum, D. H. Characterization of Lanthanide High Entropy Oxides Using Transmission Electron Microscopy. Mater. Lett. 2022, 324, 132628  DOI: 10.1016/j.matlet.2022.132628
  27. 27
    Hussien, A. G.; Damaskinos, C.; Dabbawala, A. A.; Anjum, D. H.; Vasiliades, M.; Khaleel, M.; Wehbe, N.; Efstathiou, A. M.; Polychronopoulou, K. Elucidating the Role of La3+/Sm3+ in the Carbon Paths of Dry Reforming of Methane Over Ni/Ce-La(Sm)-Cu-O Using Transient Kinetics and Isotopic Techniques. Appl. Catal., B 2022, 304, 121015  DOI: 10.1016/j.apcatb.2021.121015
  28. 28
    Polychronopoulou, K.; Dabbawala, A. A.; Sajjad, M.; Singh, N.; Anjum, D. H.; Baker, M. A.; Charisiou, N.; Goula, M. Hydrogen Production via Steam Reforming of Glycerol Over Ce-La-Cu-O Ternary Oxide Catalyst: An Experimental and DFT Study. Appl. Surf. Sci. 2022, 586, 152798,  DOI: 10.1016/j.apsusc.2022.152798
  29. 29
    Harfouche, M.; Abdellatief, M.; Momani, Y.; Abbadi, A.; Al Najdawi, M.; Al Zoubi, M.; Aljamal, B.; Matalgah, S.; Khan, L. U.; Lausi, A.; Paolucci, G. Emergence of the First XAFS/XRF Beamline in the Middle East: Providing Studies of Elements and Their Atomic/Electronic Structure in Pluridisciplinary Research Fields. J. Synchrotron Rad. 2022, 29, 11071113,  DOI: 10.1107/S1600577522005215
  30. 30
    Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: Data Analysis for X-ray Absorption Spectroscopy Using IFEFFIT. J. Synchrotron Rad. 2005, 12, 537541,  DOI: 10.1107/S0909049505012719
  31. 31
    Ressler, T. WinXAS: A Program for X-ray Absorption Spectroscopy Data Analysis Under MS-Windows. J. Synchrotron Rad. 1998, 5, 118122,  DOI: 10.1107/S0909049597019298
  32. 32
    IXAS: X-ray Absorption Data Library Online, https://xaslib.xrayabsorption.org/elem/ (accessed January 2023).
  33. 33
    F. W. Lytle database Online, http://ixs.iit.edu/database/data/Farrel_Lytle_data/RAW/periodic-table.html (accessed January 2023).
  34. 34
    Costa, C. N.; Christou, S. Y.; Georgiou, G.; Efstathiou, A. M. Mathematical Modeling of the Oxygen Storage Capacity Phenomenon Studied by CO Pulse Transient Experiments Over Pd/CeO2 Catalyst. J. Catal. 2003, 219, 259272,  DOI: 10.1016/S0021-9517(03)00151-9
  35. 35
    Perego, C.; Peratello, S. Experimental Methods in Catalytic Kinetics. Catal. Today 1999, 52, 133145,  DOI: 10.1016/S0920-5861(99)00071-1
  36. 36
    Vasiliades, M. A.; Harris, D.; Stephenson, H.; Boghosian, S.; Efstathiou, A. M. Novel Analysis of Transient Isothermal 18O Isotopic Exchange on Commercial CexZr1-xO2-Based OSC Materials. Top. Catal. 2019, 62, 219226,  DOI: 10.1007/s11244-018-1116-x
  37. 37
    Sadovskaya, E. M.; Bobin, A. S.; Skazka, V. V. Isotopic Transient Analysis of Oxygen Exchange Over Oxides. Chem. Eng. J. 2018, 348, 10251036,  DOI: 10.1016/j.cej.2018.05.027
  38. 38
    Skorodumova, N. V.; Simak, S. I.; Lundqvist, B. I.; Abrikosov, I. A.; Johansson, B. Quantum Origin of the Oxygen Storage Capability of Ceria. Phys. Rev. Lett. 2002, 89, 166601  DOI: 10.1103/PhysRevLett.89.166601
  39. 39
    Hinuma, Y.; Toyao, T.; Kamachi, T.; Maeno, Z.; Takakusagi, S.; Furukawa, S.; Takigawa, I.; Shmizu, K. Density Functional Theory Calculations of Oxygen Vacancy Formation and Subsequent Molecular Adsorption on Oxide Surfaces. J. Phys. Chem. C 2018, 122, 2943529444,  DOI: 10.1021/acs.jpcc.8b11279
  40. 40
    AlKhoori, A. A.; Polychronopoulou, K.; Belabbes, A.; Jaoude, M. A.; Vega, L. F.; Sebastian, V.; Hinder, S.; Baker, M. A.; Zedan, A. F. Cu, Sm co-doping effect on the CO oxidation activity of CeO2. A combined experimental and density functional study. Appl. Surf. Sci. 2020, 521, 146305,  DOI: 10.1016/j.apsusc.2020.146305
  41. 41
    Rakhmatullin, R.M.; Semashko, V.V.; Korableva, S.L.; Kiiamov, A.G.; Rodionov, A.A.; Tschaggelar, R.; van Bokhoven, J. A.; Paun, C. EPR study of ceria nanoparticles containing different concentration of Ce3+ ions. Mater. Chem. Phys. 2018, 219, 251257,  DOI: 10.1016/j.matchemphys.2018.08.028
  42. 42
    Alkhoori, A. A.; Elmutasim, O.; Dabbawala, A. A.; Vasiliades, M. A.; Petallidou, K. C.; Emwas, A.-H.; Anjum, D. H.; Singh, N.; Baker, M. A.; Charisiou, N. D.; Goula, M. A.; Efstathiou, A. M.; Polychronopoulou, K. Mechanistic Features of the CeO2-Modified Ni/Al2O3 Catalysts for the CO2Methanation Reaction: Experimental and Ab Initio Studies. ACS Applied Energy Materials 2023, 6 (16), 85508571,  DOI: 10.1021/acsaem.3c01437
  43. 43
    Agarwal, S.; Zhu, X.; Hensen, E. J.; Lefferts, L.; Mojet, B. L. Defect Chemistry of Ceria Nanorods. J. Phys. Chem. C 2014, 118, 41314142,  DOI: 10.1021/jp409989y
  44. 44
    Sarkar, A.; Loho, C.; Velasco, L.; Thomas, T.; Bhattachary, S. S.; Hahn, H.; Djenadic, R. Multicomponent equiatomic rare earth oxides with a narrow band gap and associated Praseodymium Multivalency. Dalton Trans. 2017, 46, 1216712176,  DOI: 10.1039/C7DT02077E
  45. 45
    Acharya, S. A.; Gaikwad, V.; D’Souza, S. W.; Barman, S. R. Gd/Sm Dopant-modified Oxidation State and Defect Generation in Nano-Ceria. Solid State Ionics 2014, 260, 2129,  DOI: 10.1016/j.ssi.2014.03.008
  46. 46
    Filtschew, A.; Hofmann, K.; Hess, C. Ceria and Its Defect Structure: New Insights From a Combined Spectroscopic Approach. J. Phys. Chem. C 2016, 120, 66946703,  DOI: 10.1021/acs.jpcc.6b00959
  47. 47
    Andriopoulou, C.; Trimpalis, A.; Petallidou, K. C.; Sgoura, A.; Efstathiou, A. M.; Boghosian, S. Structural and Redox Properties of Ce1–xZrxO2−δ and Ce0. 8Zr0.15RE0.05O2−δ (RE: La, Nd, Pr, Y) Solids Studied by High Temperature in Situ Raman Spectroscopy. J. Phys. Chem. C 2017, 121, 79317943,  DOI: 10.1021/acs.jpcc.7b00515
  48. 48
    Lee, Y.; He, G.; Akey, A. J.; Si, R.; Stephanopoulos, M. F.; Herman, I. P. Raman Analysis of Mode Softening in Nanoparticle CeO2−δ and Au-CeO2−δ During CO Oxidation. J. Am. Chem. Soc. 2011, 133, 1295212955,  DOI: 10.1021/ja204479j
  49. 49
    Wu, Z.; Li, M.; Howe, J.; Meyer, H. M.; Overbury, S. Probing Defect Sites on CeO2 Nanocrystals with Well-defined Surface Planes by Raman Spectroscopy and O2 Adsorption. Langmuir 2010, 26, 1659516606,  DOI: 10.1021/la101723w
  50. 50
    Sartoretti, E.; Novara, C.; Giorgis, F.; Piumetti, M.; Bensaid, S.; Russo, N.; Fino, D. In Situ Raman Analyses of the Soot Oxidation Reaction over Nanostructured Ceria-based Catalysts. Sci. Rep. 2019, 9, 3875,  DOI: 10.1038/s41598-019-39105-5
  51. 51
    Purans, J.; Menushenkov, A. P.; Besedin, S. P.; Ivanov, A. A.; Minkov, V. S.; Pudza, I.; Kuzmin, A.; Klementiev, K. V.; Pascarelli, S.; Mathon, O.; Rosa, A. D.; Irifune, T.; Eremets, M. I. Local electronic structure rearrangements and strong anharmonicity in YH3 under pressures up to 180 GPA. Nat. Commun. 2021, 12, 1765,  DOI: 10.1038/s41467-021-21991-x
  52. 52
    Satsuma, A.; Yanagihara, M.; Ohyama, J.; Shimizua, K. Oxidation of CO over Ru/Ceria prepared by self-dispersion of Ru metal powder into nano-sized particle. Catal. Today 2013, 201, 6267,  DOI: 10.1016/j.cattod.2012.03.048
  53. 53
    Loridant, S. Raman spectroscopy as a powerful tool to characterize ceria-based catalysts. Catal. Today 2021, 373, 98111,  DOI: 10.1016/j.cattod.2020.03.044
  54. 54
    Wang, H.; Li, X.; Ruan, Q.; Tang, J. Ru and Ruoxdecorated carbon nitride for efficient ammonia photosynthesis. Nanoscale 2020, 12, 1232912335,  DOI: 10.1039/D0NR02527E
  55. 55
    Gu, Y. 1,2-dichloroethane deep oxidation over bifunctional Ru/CExAly Catalysts. ACS Omega 2018, 3, 84608470,  DOI: 10.1021/acsomega.8b00592
  56. 56
    Siakavelas, G. I. Highly selective and stable nickel catalysts supported on Ceria promoted with SM2O3, PR2O3 and mgo for the CO2Methanation reaction. Applied Catalysis B: Environmental 2021, 282, 119562,  DOI: 10.1016/j.apcatb.2020.119562
  57. 57
    Polychronopoulou, K. Decoupling the chemical and mechanical strain effect on steering the CO2 activation over CeO2-based oxides: An experimental and DFT approach. ACS Appl. Mater. Interfaces 2022, 14, 3309433119,  DOI: 10.1021/acsami.2c05714
  58. 58
    Younis, A.; Shirsath, S. E.; Shabbir, B.; Li, S. Controllable dynamics of oxygen vacancies through extrinsic doping for superior catalytic activities. Nanoscale 2018, 10, 1857618585,  DOI: 10.1039/C8NR03801E
  59. 59
    Holder, C. F.; Schaak, R. E. Tutorial on Powder X-ray Diffraction for Characterizing Nanoscale Materials. ACS Nano 2019, 13, 73597365,  DOI: 10.1021/acsnano.9b05157
  60. 60
    Schmitt, R.; Nenning, A.; Kraynis, O.; Korobko, R.; Frenkel, A. I.; Lubomirsky, I.; Haile, S. M.; Rupp, J. L. M. A Review of Defect Structure and Chemistry in Ceria and Its Solid Solutions. J. Chem. Soc. Rev. 2020, 49, 554592,  DOI: 10.1039/C9CS00588A
  61. 61
    Jacobsen, J.; Dahl, S.; Hansen, P. L.; Tornqvist, E.; Jensen, L.; Topsoe, H.; Prip, D. V.; Moenshaug, P. B.; Chorkendorff, I. Structure Sensitivity of Supported Ruthenium Catalysts for Ammonia Synthesis. J. Mol. Catal. A: Chem. 2000, 163, 1926,  DOI: 10.1016/S1381-1169(00)00396-4
  62. 62
    Lingling, L.; Ya-Fei, J.; Tianhua, Z.; Hongfang, C.; Yanliang, Z.; Bingyu, L.; Xingyi, L.; Ying, Z.; Lirong, Z.; Xiuyun, W.; Cong-Qiao, X.; Chak-tong, A.; Lilong, J.; Jun, L. Size sensitivity of supported Ru catalysts for ammonia synthesis: From nanoparticles to subnanometric clusters and atomic clusters. Chem 2022, 8, 749768,  DOI: 10.1016/j.chempr.2021.11.008
  63. 63
    Raróg-Pilecka, W.; Miśkiewicz, E.; Szmigiel, D.; Kowalczyk, Z. Structure sensitivity of ammonia synthesis over promoted ruthenium catalysts supported on graphitised carbon. J. Catal. 2005, 231, 1119,  DOI: 10.1016/j.jcat.2004.12.005
  64. 64
    Kim, Y. K.; Morgan, G. A.; Yates, J. T. Site-specific dissociation of N-2 on the stepped Ru(109). Surf. Sci. 2005, 14, 598
  65. 65
    Lin, H.; Liu, J. X.; Fan, H.-J.; Li, W.-X. Crystallographic and morphological sensitivity of N2 activation over ruthenium. Chin. J. Chem. Phys. 2021, 34, 263272,  DOI: 10.1063/1674-0068/cjcp2009171
  66. 66
    Jiang, Rongzhong; Tran, Dat T.; Li, Jiangtian; Chu, D. Ru@RuO2 Core-Shell Nanorods: A Highly Active and Stable Bifunctional Catalyst for Oxygen Evolution and Hydrogen Evolution Reactions. Energy Environ. Mater. 2019, 2, 201208,  DOI: 10.1002/eem2.12031
  67. 67
    Chandra, D.; Inoue, Y.; Sasase, M.; Kitano, M.; Bhaumik, A.; Kamata, K.; Hosono, H.; Hara, M. Chem. Sci. 2018, 9, 59495956,  DOI: 10.1039/C8SC01197D
  68. 68
    Stevens, C. A. C.; Lambie, S. G.; Ruffman, C.; Skulason, E.; Garden, A. L. Geometric and Electronic Effects Contributing to N2 Dissociation Barriers on a Range of Active Sites on Ru Nanoparticles. J. Phys. Chem. C 2019, 123, 3045830466,  DOI: 10.1021/acs.jpcc.9b09563
  69. 69
    Tang, C.; Qiao, S.-Z. How to explore ambient electrocatalytic nitrogen reduction reliably and insightfully. Chem. Soc. Rev. 2019, 48, 31663180,  DOI: 10.1039/C9CS00280D
  70. 70
    Singh, S.; Mohammed, A. K.; AlHammadi, A. A.; Shetty, D.; Polychronopoulou, K. Hypes and hopes on the materials development strategies to produce ammonia at mild conditions. Int. J. Hydrogen Energy 2023, 48, 3470034739,  DOI: 10.1016/j.ijhydene.2023.05.206
  71. 71
    Zhou, Y.; Wang, J.; Liang, L.; Sai, Q.; Ni, J.; Au, C. T.; Lin, X.; Wang, X.; Zheng, Y.; Zheng, L.; Jiang, L. Unraveling the size-dependent effect of RU-based catalysts on ammonia synthesis at mild conditions. J. Catal. 2021, 404, 501511,  DOI: 10.1016/j.jcat.2021.10.024
  72. 72
    Peng, X.; Chen, X.; Zhou, Y.; Sun, F.; Zhang, T.; Zheng, L.; Jiang, L.; Wang, X. Size-dependent activity of supported RU catalysts for ammonia synthesis at mild conditions. J. Catal. 2022, 408, 98108,  DOI: 10.1016/j.jcat.2022.02.024
  73. 73
    Hinrichsen, O.; Rosowski, F.; Hornung, A.; Muhler, M.; Ertl, G. The kinetics of ammonia synthesis over RU-based catalysts. J. Catal. 1997, 165, 3344,  DOI: 10.1006/jcat.1997.1447
  74. 74
    Qiu, J.-Z.; Hu, J.; Lan, J.; Wang, L.-F.; Fu, G.; Xiao, R.; Ge, B.; Jiang, J. Pure siliceous zeolite-supported RU single-atom active sites for ammonia synthesis. Chem. Mater. 2019, 31, 94139421,  DOI: 10.1021/acs.chemmater.9b03099
  75. 75
    Wang, X.; Li, L.; Fang, Z.; Zhang, Y.; Ni, J.; Lin, B.; Zheng, L.; Au, C. T.; Jiang, L. Atomically dispersed Ru catalyst for low-temperature nitrogen activation to ammonia via an associative mechanism. ACS Catal. 2020, 10, 95049514,  DOI: 10.1021/acscatal.0c00549
  76. 76
    Yao, Y.; Zhu, S.; Wang, H.; Li, H.; Shao, M. A spectroscopic study on the nitrogen electrochemical reduction reaction on Gold and platinum surfaces. J. Am. Chem. Soc. 2018, 140, 14961501,  DOI: 10.1021/jacs.7b12101
  77. 77
    Schmitt, R.; Nenning, A.; Kraynis, O.; Korobko, R.; Frenkel, A. I.; Lubomirsky, I.; Haile, S. M.; Rupp, J. L. M. A review of defect structure and chemistry in Ceria and its solid solutions. Chem. Soc. Rev. 2020, 49, 554592,  DOI: 10.1039/C9CS00588A
  78. 78
    Cheng, S.; Gao, Y.-J.; Yan, Y.-L.; Gao, X.; Zhang, S.-H.; Zhuang, G.-L.; Deng, S.-W.; Wei, Z.-Z.; Zhong, X.; Wang, J.-G. Oxygen vacancy enhancing mechanism of nitrogen reduction reaction property in Ru/TiO2. Journal of Energy Chemistry 2019, 39, 144151,  DOI: 10.1016/j.jechem.2019.01.020

Cited By

Click to copy section linkSection link copied!

This article has not yet been cited by other publications.

Open PDF

ACS Applied Materials & Interfaces

Cite this: ACS Appl. Mater. Interfaces 2024, 16, 18, 23038–23053
Click to copy citationCitation copied!
https://doi.org/10.1021/acsami.3c16521
Published April 29, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

1887

Altmetric

-

Citations

-
Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Unit cells of (a) pure CeO2, (b) binary CeGdO, (c) ternary CeLaGdO, and (d) quinary CeLaPrSmGdO.

    Figure 2

    Figure 2. (a) Unit cell of CeO2 along (110) surface with vacuum. Bulk region, sublayer, and top-layers are shown. (b) Evolution of oxygen vacancy formation energy in pure CeO2, binary systems (CeGdO, CePrO), ternary systems (CeLaGdO, CeLaPrO), and quinary (CeLaPrSmGdO) high entropy system.

    Figure 3

    Figure 3. (A) XRD diffractograms; (B–C) EPR spectra collected at 100 K; (D) Raman spectra collected over the Ru/CePrO catalyst; (E) Raman spectra collected over the Ru/HEO catalyst; (F) H2-temperature-programmed desorption (H2-TPD) profiles obtained over all the Ru catalysts of the present study.

    Figure 4

    Figure 4. (A) Ru 3p, (B) Ce 3d, and (C) O 1s XPS core-level spectra of the prepared catalysts.

    Figure 5

    Figure 5. (A, B) Linear combination fitting of the XANES spectra collected at the LIII edge of Ce in the HEO sample (left) and CePrO sample (right) to determine all the oxidation states of Ce in the samples; (C) Fourier transforms with their respective real part of the k3 weighted EXAFS spectra collected at the Ce LIII edge in CLPSG-HEO and CePr samples.

    Figure 6

    Figure 6. (A) HRTEM images along with Fast Fourier Transform (FFT) pattern; (B, C) facets analysis of the Ru/CePrO catalyst.

    Figure 7

    Figure 7. (A–E) STEM-HAADF RGB analysis and (F) selected area electron diffraction (SAED) over the Ru/CePrO catalyst.

    Figure 8

    Figure 8. (A–C) HRTEM images of the Ru/HEO catalyst with emphasis on Ru particles at different areas (B) and (C).

    Figure 9

    Figure 9. (A) HRTEM images of the Ru/HEO catalyst along with Fast Fourier Transform (FFT) pattern and (B–D) facets analysis.

    Figure 10

    Figure 10. (A–H) STEM-HAADF RGB analysis and (I) selected area electron diffraction (SAED) over the Ru/HEO catalyst.

    Figure 11

    Figure 11. Catalytic production of NH3 (A) over Ru-based catalysts of this study in the 10–40 bar pressure range and 400 °C; (B) over the Ru/HEO catalyst at 10 bar, 400 °C following two different activation conditions (650 vs 800 °C).

    Figure 12

    Figure 12. (A) Transient rates of 16O2 (μmol g–1 s–1) consumed during transient isothermal oxidation (TIO) at 650 °C, following the different pretreatment conditions, over HEO solid; and (B) amounts of O2 consumed (μmol g–1) during TIO, over both solids.

    Figure 13

    Figure 13. (A) Transient concentration (mol %) response curves of Kr, 16O2, 16O18O, and 18O2 gaseous species recorded during the 18O2-TIIE experiment at 650 °C, over the calcined HEO; (B) αg(18) descriptor as a function of time and pretreatment procedure, estimated during 18O2-TIIE over HEO; and (C) αg(18) descriptor obtained over the both reduced at 650 °C solids.

    Figure 14

    Figure 14. Top and side views of N2 Adsorption configurations on Ru4/reduced HEO surface: (a) in absence of hydrogen and (b) in the presence of hydrogen. Gd is shown in purple, La in dark green, Pr in yellow, Sm in pink, Ce in light green, Ru in orange, N in white, H in blue, and O in red.

    Figure 15

    Figure 15. Top and side views of N Adsorption configurations on reduced quinary CeLaPrSmGdO (111) surface: (a) in absence of hydrogen and (b) in the presence of hydrogen. Gd is shown in purple, La in dark green, Pr in yellow, Ce in light green, Sm in pink, H in blue, N in white, and O in red.

  • References


    This article references 78 other publications.

    1. 1
      Mehta, P.; Barboun, P.; Herrera, F. A.; Kim, J.; Rumbach, P.; Go, D. B.; Hicks, J. C.; Scheider, W. F. Overcoming Ammonia Synthesis Scaling Relations with Plasma-enabled Catalysis. Nat. Catal. 2018, 1, 269275,  DOI: 10.1038/s41929-018-0045-1
    2. 2
      Guo, C.; Ran, J.; Vasileff, A.; Qiao, S. Z. Rational Design of Electrocatalysts and Photo(electro)catalysts for Nitrogen Reduction to Ammonia (NH3) under Ambient Conditions. Energy Environ. Sci. 2018, 11, 4556,  DOI: 10.1039/C7EE02220D
    3. 3
      Kordali, V.; Kyriacou, G.; Lambrou, Ch. Electrochemical Synthesis of Ammonia at Atmospheric Pressure and Low Temperature in a Solid Polymer Electrolyte Cell. Chem. Commun. 2000, 16731674,  DOI: 10.1039/b004885m
    4. 4
      Tao, H.; Choi, C.; Ding, L. X.; Jiang, Z.; Han, Z.; Jia, M.; Fan, Q.; Gao, Y.; Wang, H.; Robertson, A. W.; Hong, S.; Jung, Y.; Liu, S.; Sun, Z. Nitrogen Fixation by Ru Single-Atom Electrocatalytic Reduction. Chem. 2019, 5, 204214,  DOI: 10.1016/j.chempr.2018.10.007
    5. 5
      Yao, H. C.; Yao, Y. F. Ceria in Automotive Exhaust Catalysts: I. Oxygen Storage. J. Catal. 1984, 86, 254265,  DOI: 10.1016/0021-9517(84)90371-3
    6. 6
      Petallidou, K. C.; Polychronopoulou, K.; Boghosian, S.; Rodriguez, S. G.; Efstathiou, A. M. Water–Gas Shift Reaction on Pt/Ce1–xTixO2−δ: The Effect of Ce/Ti Ratio. J. Phys. Chem. C 2013, 117, 2546725477,  DOI: 10.1021/jp406059h
    7. 7
      Dou, B.; Zhao, L.; Zhang, H.; Wu, K.; Zhang, H. Renewable hydrogen production from chemical looping steam reforming of biodiesel byproduct glycerol by mesoporous oxygen carriers. Chemical Engineering Journal 2021, 416, 127612,  DOI: 10.1016/j.cej.2020.127612
    8. 8
      Kim, Y. M.; Lohsoontorn, P. K.; Bae, J. Effect of Unsintered Gadolinium-doped Ceria Buffer Layer on Performance of Metal-supported Solid Oxide Fuel Cells Using Unsintered Barium Strontium Cobalt Ferrite Cathode. J. Power Sources 2010, 195, 64206427,  DOI: 10.1016/j.jpowsour.2010.03.095
    9. 9
      Sarkar, A.; Velasco, L.; Wang, D.; Wang, Q.; Talasila, G.; Biasi, L. D.; Kubel, C.; Brezesinski, T.; Bhattacharya, S. S.; Hahn, H.; Breitung, B. High Entropy Oxides for Reversible Energy Storage. Nat. Commun. 2018, 9, 3400,  DOI: 10.1038/s41467-018-05774-5
    10. 10
      Albedwawi, S. H.; Aljaberi, A.; Haidemenopoulos, G. N.; Polychronopoulou, K. High Entropy Oxides-Exploring a Paradigm of Promising Catalysts: A Review. Materials & Design 2021, 202, 109534,  DOI: 10.1016/j.matdes.2021.109534
    11. 11
      Chae, S.; Williams, L.; Lee, J.; Heron, J. T.; Kioupakis, E. Effects of Local Compositional and Structural Disorder on Vacancy Formation in Entropy-stabilized Oxides from First-principles. npj Comput. Mater. 2022, 8, 95,  DOI: 10.1038/s41524-022-00780-0
    12. 12
      Ye, S.; Zhu, J.; Zhu, S.; Zhao, Y.; Li, M.; Huang, Z.; Wang, H.; He, J. Design strategies for perovskite-type high-entropy oxides with applications in optics. ACS Appl. Mater. Interfaces 2023, 15, 4747547486,  DOI: 10.1021/acsami.3c09447
    13. 13
      Sarkar, A.; Kruk, R.; Hahn, H. Magnetic properties of high entropy oxides. Dalton Transactions 2021, 50, 19731982,  DOI: 10.1039/D0DT04154H
    14. 14
      Sarkar, A.; Wang, Q.; Schiele, A.; Chellali, M. R.; Bhattacharya, S. S.; Wang, D.; Brezesinski, T.; Hahn, H.; Velasco, L.; Breitung, B. High-entropy oxides: Fundamental aspects and electrochemical properties. Adv. Mater. 2019, 31, e1806236  DOI: 10.1002/adma.201806236
    15. 15
      Brahlek, M.; Gazda, M.; Keppens, V.; Mazza, A. R.; McCormack, S. J.; Gryn, A. M.; Musico, B.; Page, K.; Rost, C. M.; Sinnott, S. B.; Toher, C.; Ward, T. Z.; Yamamoto, A. What is in a Name: Defining “High Entropy” Oxides. APL Mater. 2022, 10, 110902  DOI: 10.1063/5.0122727
    16. 16
      Kresse, G.; Furthmuller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 1116911186,  DOI: 10.1103/PhysRevB.54.11169
    17. 17
      Kresse, G.; Furthmuller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 1550,  DOI: 10.1016/0927-0256(96)00008-0
    18. 18
      Pedrew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
    19. 19
      Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953,  DOI: 10.1103/PhysRevB.50.17953
    20. 20
      Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-wave Method. Phys. Rev. B 1999, 59, 1758,  DOI: 10.1103/PhysRevB.59.1758
    21. 21
      Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.; Sutton, A. P. Electron-energy-loss Spectra and the Structural Stability of Nickel Oxide: An LSDA+U Study. Phys. Rev. B 1998, 57, 1505,  DOI: 10.1103/PhysRevB.57.1505
    22. 22
      Zunger, A.; Wei, S. H.; Ferreira, L. G.; Bernard, J. E. Special Quasi-random Structures. Phys. Rev. Lett. 1990, 65, 353
    23. 23
      Walle, A. V.; Tiwary, P.; Jong, M. D.; Olmsted, D. L.; Asta, M.; Dick, A.; Shin, D.; Wang, Y.; Chen, L. Q.; Liu, Z. K. Efficient Stochastic Generation of Special Quasi-random Structures. Calphad 2013, 42, 1318,  DOI: 10.1016/j.calphad.2013.06.006
    24. 24
      Ghosh, G.; Walle, A. V.; Asta, M. First-principles Calculations of the Structural and Thermodynamic Properties of bcc, fcc and hcp Solid Solutions in the Al–TM (TM = Ti, Zr and Hf) Systems: A Comparison of Cluster Expansion and Supercell Methods. Acta Mater. 2008, 56, 3202,  DOI: 10.1016/j.actamat.2008.03.006
    25. 25
      Qu, P. F.; Wang, G. C. Theoretical Insight into the Strong Size-dependence of Dry Reforming of Methane Over Ru/CeO2. J. CO2 Util. 2022, 65, 102221,  DOI: 10.1016/j.jcou.2022.102221
    26. 26
      Zafar, H.; Albedwawi, S.; Polychronopoulou, K.; Anjum, D. H. Characterization of Lanthanide High Entropy Oxides Using Transmission Electron Microscopy. Mater. Lett. 2022, 324, 132628  DOI: 10.1016/j.matlet.2022.132628
    27. 27
      Hussien, A. G.; Damaskinos, C.; Dabbawala, A. A.; Anjum, D. H.; Vasiliades, M.; Khaleel, M.; Wehbe, N.; Efstathiou, A. M.; Polychronopoulou, K. Elucidating the Role of La3+/Sm3+ in the Carbon Paths of Dry Reforming of Methane Over Ni/Ce-La(Sm)-Cu-O Using Transient Kinetics and Isotopic Techniques. Appl. Catal., B 2022, 304, 121015  DOI: 10.1016/j.apcatb.2021.121015
    28. 28
      Polychronopoulou, K.; Dabbawala, A. A.; Sajjad, M.; Singh, N.; Anjum, D. H.; Baker, M. A.; Charisiou, N.; Goula, M. Hydrogen Production via Steam Reforming of Glycerol Over Ce-La-Cu-O Ternary Oxide Catalyst: An Experimental and DFT Study. Appl. Surf. Sci. 2022, 586, 152798,  DOI: 10.1016/j.apsusc.2022.152798
    29. 29
      Harfouche, M.; Abdellatief, M.; Momani, Y.; Abbadi, A.; Al Najdawi, M.; Al Zoubi, M.; Aljamal, B.; Matalgah, S.; Khan, L. U.; Lausi, A.; Paolucci, G. Emergence of the First XAFS/XRF Beamline in the Middle East: Providing Studies of Elements and Their Atomic/Electronic Structure in Pluridisciplinary Research Fields. J. Synchrotron Rad. 2022, 29, 11071113,  DOI: 10.1107/S1600577522005215
    30. 30
      Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: Data Analysis for X-ray Absorption Spectroscopy Using IFEFFIT. J. Synchrotron Rad. 2005, 12, 537541,  DOI: 10.1107/S0909049505012719
    31. 31
      Ressler, T. WinXAS: A Program for X-ray Absorption Spectroscopy Data Analysis Under MS-Windows. J. Synchrotron Rad. 1998, 5, 118122,  DOI: 10.1107/S0909049597019298
    32. 32
      IXAS: X-ray Absorption Data Library Online, https://xaslib.xrayabsorption.org/elem/ (accessed January 2023).
    33. 33
      F. W. Lytle database Online, http://ixs.iit.edu/database/data/Farrel_Lytle_data/RAW/periodic-table.html (accessed January 2023).
    34. 34
      Costa, C. N.; Christou, S. Y.; Georgiou, G.; Efstathiou, A. M. Mathematical Modeling of the Oxygen Storage Capacity Phenomenon Studied by CO Pulse Transient Experiments Over Pd/CeO2 Catalyst. J. Catal. 2003, 219, 259272,  DOI: 10.1016/S0021-9517(03)00151-9
    35. 35
      Perego, C.; Peratello, S. Experimental Methods in Catalytic Kinetics. Catal. Today 1999, 52, 133145,  DOI: 10.1016/S0920-5861(99)00071-1
    36. 36
      Vasiliades, M. A.; Harris, D.; Stephenson, H.; Boghosian, S.; Efstathiou, A. M. Novel Analysis of Transient Isothermal 18O Isotopic Exchange on Commercial CexZr1-xO2-Based OSC Materials. Top. Catal. 2019, 62, 219226,  DOI: 10.1007/s11244-018-1116-x
    37. 37
      Sadovskaya, E. M.; Bobin, A. S.; Skazka, V. V. Isotopic Transient Analysis of Oxygen Exchange Over Oxides. Chem. Eng. J. 2018, 348, 10251036,  DOI: 10.1016/j.cej.2018.05.027
    38. 38
      Skorodumova, N. V.; Simak, S. I.; Lundqvist, B. I.; Abrikosov, I. A.; Johansson, B. Quantum Origin of the Oxygen Storage Capability of Ceria. Phys. Rev. Lett. 2002, 89, 166601  DOI: 10.1103/PhysRevLett.89.166601
    39. 39
      Hinuma, Y.; Toyao, T.; Kamachi, T.; Maeno, Z.; Takakusagi, S.; Furukawa, S.; Takigawa, I.; Shmizu, K. Density Functional Theory Calculations of Oxygen Vacancy Formation and Subsequent Molecular Adsorption on Oxide Surfaces. J. Phys. Chem. C 2018, 122, 2943529444,  DOI: 10.1021/acs.jpcc.8b11279
    40. 40
      AlKhoori, A. A.; Polychronopoulou, K.; Belabbes, A.; Jaoude, M. A.; Vega, L. F.; Sebastian, V.; Hinder, S.; Baker, M. A.; Zedan, A. F. Cu, Sm co-doping effect on the CO oxidation activity of CeO2. A combined experimental and density functional study. Appl. Surf. Sci. 2020, 521, 146305,  DOI: 10.1016/j.apsusc.2020.146305
    41. 41
      Rakhmatullin, R.M.; Semashko, V.V.; Korableva, S.L.; Kiiamov, A.G.; Rodionov, A.A.; Tschaggelar, R.; van Bokhoven, J. A.; Paun, C. EPR study of ceria nanoparticles containing different concentration of Ce3+ ions. Mater. Chem. Phys. 2018, 219, 251257,  DOI: 10.1016/j.matchemphys.2018.08.028
    42. 42
      Alkhoori, A. A.; Elmutasim, O.; Dabbawala, A. A.; Vasiliades, M. A.; Petallidou, K. C.; Emwas, A.-H.; Anjum, D. H.; Singh, N.; Baker, M. A.; Charisiou, N. D.; Goula, M. A.; Efstathiou, A. M.; Polychronopoulou, K. Mechanistic Features of the CeO2-Modified Ni/Al2O3 Catalysts for the CO2Methanation Reaction: Experimental and Ab Initio Studies. ACS Applied Energy Materials 2023, 6 (16), 85508571,  DOI: 10.1021/acsaem.3c01437
    43. 43
      Agarwal, S.; Zhu, X.; Hensen, E. J.; Lefferts, L.; Mojet, B. L. Defect Chemistry of Ceria Nanorods. J. Phys. Chem. C 2014, 118, 41314142,  DOI: 10.1021/jp409989y
    44. 44
      Sarkar, A.; Loho, C.; Velasco, L.; Thomas, T.; Bhattachary, S. S.; Hahn, H.; Djenadic, R. Multicomponent equiatomic rare earth oxides with a narrow band gap and associated Praseodymium Multivalency. Dalton Trans. 2017, 46, 1216712176,  DOI: 10.1039/C7DT02077E
    45. 45
      Acharya, S. A.; Gaikwad, V.; D’Souza, S. W.; Barman, S. R. Gd/Sm Dopant-modified Oxidation State and Defect Generation in Nano-Ceria. Solid State Ionics 2014, 260, 2129,  DOI: 10.1016/j.ssi.2014.03.008
    46. 46
      Filtschew, A.; Hofmann, K.; Hess, C. Ceria and Its Defect Structure: New Insights From a Combined Spectroscopic Approach. J. Phys. Chem. C 2016, 120, 66946703,  DOI: 10.1021/acs.jpcc.6b00959
    47. 47
      Andriopoulou, C.; Trimpalis, A.; Petallidou, K. C.; Sgoura, A.; Efstathiou, A. M.; Boghosian, S. Structural and Redox Properties of Ce1–xZrxO2−δ and Ce0. 8Zr0.15RE0.05O2−δ (RE: La, Nd, Pr, Y) Solids Studied by High Temperature in Situ Raman Spectroscopy. J. Phys. Chem. C 2017, 121, 79317943,  DOI: 10.1021/acs.jpcc.7b00515
    48. 48
      Lee, Y.; He, G.; Akey, A. J.; Si, R.; Stephanopoulos, M. F.; Herman, I. P. Raman Analysis of Mode Softening in Nanoparticle CeO2−δ and Au-CeO2−δ During CO Oxidation. J. Am. Chem. Soc. 2011, 133, 1295212955,  DOI: 10.1021/ja204479j
    49. 49
      Wu, Z.; Li, M.; Howe, J.; Meyer, H. M.; Overbury, S. Probing Defect Sites on CeO2 Nanocrystals with Well-defined Surface Planes by Raman Spectroscopy and O2 Adsorption. Langmuir 2010, 26, 1659516606,  DOI: 10.1021/la101723w
    50. 50
      Sartoretti, E.; Novara, C.; Giorgis, F.; Piumetti, M.; Bensaid, S.; Russo, N.; Fino, D. In Situ Raman Analyses of the Soot Oxidation Reaction over Nanostructured Ceria-based Catalysts. Sci. Rep. 2019, 9, 3875,  DOI: 10.1038/s41598-019-39105-5
    51. 51
      Purans, J.; Menushenkov, A. P.; Besedin, S. P.; Ivanov, A. A.; Minkov, V. S.; Pudza, I.; Kuzmin, A.; Klementiev, K. V.; Pascarelli, S.; Mathon, O.; Rosa, A. D.; Irifune, T.; Eremets, M. I. Local electronic structure rearrangements and strong anharmonicity in YH3 under pressures up to 180 GPA. Nat. Commun. 2021, 12, 1765,  DOI: 10.1038/s41467-021-21991-x
    52. 52
      Satsuma, A.; Yanagihara, M.; Ohyama, J.; Shimizua, K. Oxidation of CO over Ru/Ceria prepared by self-dispersion of Ru metal powder into nano-sized particle. Catal. Today 2013, 201, 6267,  DOI: 10.1016/j.cattod.2012.03.048
    53. 53
      Loridant, S. Raman spectroscopy as a powerful tool to characterize ceria-based catalysts. Catal. Today 2021, 373, 98111,  DOI: 10.1016/j.cattod.2020.03.044
    54. 54
      Wang, H.; Li, X.; Ruan, Q.; Tang, J. Ru and Ruoxdecorated carbon nitride for efficient ammonia photosynthesis. Nanoscale 2020, 12, 1232912335,  DOI: 10.1039/D0NR02527E
    55. 55
      Gu, Y. 1,2-dichloroethane deep oxidation over bifunctional Ru/CExAly Catalysts. ACS Omega 2018, 3, 84608470,  DOI: 10.1021/acsomega.8b00592
    56. 56
      Siakavelas, G. I. Highly selective and stable nickel catalysts supported on Ceria promoted with SM2O3, PR2O3 and mgo for the CO2Methanation reaction. Applied Catalysis B: Environmental 2021, 282, 119562,  DOI: 10.1016/j.apcatb.2020.119562
    57. 57
      Polychronopoulou, K. Decoupling the chemical and mechanical strain effect on steering the CO2 activation over CeO2-based oxides: An experimental and DFT approach. ACS Appl. Mater. Interfaces 2022, 14, 3309433119,  DOI: 10.1021/acsami.2c05714
    58. 58
      Younis, A.; Shirsath, S. E.; Shabbir, B.; Li, S. Controllable dynamics of oxygen vacancies through extrinsic doping for superior catalytic activities. Nanoscale 2018, 10, 1857618585,  DOI: 10.1039/C8NR03801E
    59. 59
      Holder, C. F.; Schaak, R. E. Tutorial on Powder X-ray Diffraction for Characterizing Nanoscale Materials. ACS Nano 2019, 13, 73597365,  DOI: 10.1021/acsnano.9b05157
    60. 60
      Schmitt, R.; Nenning, A.; Kraynis, O.; Korobko, R.; Frenkel, A. I.; Lubomirsky, I.; Haile, S. M.; Rupp, J. L. M. A Review of Defect Structure and Chemistry in Ceria and Its Solid Solutions. J. Chem. Soc. Rev. 2020, 49, 554592,  DOI: 10.1039/C9CS00588A
    61. 61
      Jacobsen, J.; Dahl, S.; Hansen, P. L.; Tornqvist, E.; Jensen, L.; Topsoe, H.; Prip, D. V.; Moenshaug, P. B.; Chorkendorff, I. Structure Sensitivity of Supported Ruthenium Catalysts for Ammonia Synthesis. J. Mol. Catal. A: Chem. 2000, 163, 1926,  DOI: 10.1016/S1381-1169(00)00396-4
    62. 62
      Lingling, L.; Ya-Fei, J.; Tianhua, Z.; Hongfang, C.; Yanliang, Z.; Bingyu, L.; Xingyi, L.; Ying, Z.; Lirong, Z.; Xiuyun, W.; Cong-Qiao, X.; Chak-tong, A.; Lilong, J.; Jun, L. Size sensitivity of supported Ru catalysts for ammonia synthesis: From nanoparticles to subnanometric clusters and atomic clusters. Chem 2022, 8, 749768,  DOI: 10.1016/j.chempr.2021.11.008
    63. 63
      Raróg-Pilecka, W.; Miśkiewicz, E.; Szmigiel, D.; Kowalczyk, Z. Structure sensitivity of ammonia synthesis over promoted ruthenium catalysts supported on graphitised carbon. J. Catal. 2005, 231, 1119,  DOI: 10.1016/j.jcat.2004.12.005
    64. 64
      Kim, Y. K.; Morgan, G. A.; Yates, J. T. Site-specific dissociation of N-2 on the stepped Ru(109). Surf. Sci. 2005, 14, 598
    65. 65
      Lin, H.; Liu, J. X.; Fan, H.-J.; Li, W.-X. Crystallographic and morphological sensitivity of N2 activation over ruthenium. Chin. J. Chem. Phys. 2021, 34, 263272,  DOI: 10.1063/1674-0068/cjcp2009171
    66. 66
      Jiang, Rongzhong; Tran, Dat T.; Li, Jiangtian; Chu, D. Ru@RuO2 Core-Shell Nanorods: A Highly Active and Stable Bifunctional Catalyst for Oxygen Evolution and Hydrogen Evolution Reactions. Energy Environ. Mater. 2019, 2, 201208,  DOI: 10.1002/eem2.12031
    67. 67
      Chandra, D.; Inoue, Y.; Sasase, M.; Kitano, M.; Bhaumik, A.; Kamata, K.; Hosono, H.; Hara, M. Chem. Sci. 2018, 9, 59495956,  DOI: 10.1039/C8SC01197D
    68. 68
      Stevens, C. A. C.; Lambie, S. G.; Ruffman, C.; Skulason, E.; Garden, A. L. Geometric and Electronic Effects Contributing to N2 Dissociation Barriers on a Range of Active Sites on Ru Nanoparticles. J. Phys. Chem. C 2019, 123, 3045830466,  DOI: 10.1021/acs.jpcc.9b09563
    69. 69
      Tang, C.; Qiao, S.-Z. How to explore ambient electrocatalytic nitrogen reduction reliably and insightfully. Chem. Soc. Rev. 2019, 48, 31663180,  DOI: 10.1039/C9CS00280D
    70. 70
      Singh, S.; Mohammed, A. K.; AlHammadi, A. A.; Shetty, D.; Polychronopoulou, K. Hypes and hopes on the materials development strategies to produce ammonia at mild conditions. Int. J. Hydrogen Energy 2023, 48, 3470034739,  DOI: 10.1016/j.ijhydene.2023.05.206
    71. 71
      Zhou, Y.; Wang, J.; Liang, L.; Sai, Q.; Ni, J.; Au, C. T.; Lin, X.; Wang, X.; Zheng, Y.; Zheng, L.; Jiang, L. Unraveling the size-dependent effect of RU-based catalysts on ammonia synthesis at mild conditions. J. Catal. 2021, 404, 501511,  DOI: 10.1016/j.jcat.2021.10.024
    72. 72
      Peng, X.; Chen, X.; Zhou, Y.; Sun, F.; Zhang, T.; Zheng, L.; Jiang, L.; Wang, X. Size-dependent activity of supported RU catalysts for ammonia synthesis at mild conditions. J. Catal. 2022, 408, 98108,  DOI: 10.1016/j.jcat.2022.02.024
    73. 73
      Hinrichsen, O.; Rosowski, F.; Hornung, A.; Muhler, M.; Ertl, G. The kinetics of ammonia synthesis over RU-based catalysts. J. Catal. 1997, 165, 3344,  DOI: 10.1006/jcat.1997.1447
    74. 74
      Qiu, J.-Z.; Hu, J.; Lan, J.; Wang, L.-F.; Fu, G.; Xiao, R.; Ge, B.; Jiang, J. Pure siliceous zeolite-supported RU single-atom active sites for ammonia synthesis. Chem. Mater. 2019, 31, 94139421,  DOI: 10.1021/acs.chemmater.9b03099
    75. 75
      Wang, X.; Li, L.; Fang, Z.; Zhang, Y.; Ni, J.; Lin, B.; Zheng, L.; Au, C. T.; Jiang, L. Atomically dispersed Ru catalyst for low-temperature nitrogen activation to ammonia via an associative mechanism. ACS Catal. 2020, 10, 95049514,  DOI: 10.1021/acscatal.0c00549
    76. 76
      Yao, Y.; Zhu, S.; Wang, H.; Li, H.; Shao, M. A spectroscopic study on the nitrogen electrochemical reduction reaction on Gold and platinum surfaces. J. Am. Chem. Soc. 2018, 140, 14961501,  DOI: 10.1021/jacs.7b12101
    77. 77
      Schmitt, R.; Nenning, A.; Kraynis, O.; Korobko, R.; Frenkel, A. I.; Lubomirsky, I.; Haile, S. M.; Rupp, J. L. M. A review of defect structure and chemistry in Ceria and its solid solutions. Chem. Soc. Rev. 2020, 49, 554592,  DOI: 10.1039/C9CS00588A
    78. 78
      Cheng, S.; Gao, Y.-J.; Yan, Y.-L.; Gao, X.; Zhang, S.-H.; Zhuang, G.-L.; Deng, S.-W.; Wei, Z.-Z.; Zhong, X.; Wang, J.-G. Oxygen vacancy enhancing mechanism of nitrogen reduction reaction property in Ru/TiO2. Journal of Energy Chemistry 2019, 39, 144151,  DOI: 10.1016/j.jechem.2019.01.020
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsami.3c16521.

    • Experimental protocol for catalysts characterization that was followed in Raman spectroscopy; X-ray photoelectron spectroscopy; high-resolution transmission electron microscopy (HRTEM), electron paramagnetic resonance (EPR), and H2 chemisorption (H2-temperature-programmed desorption, TPD); textural analysis results as obtained after N2 adsorption at 77 K; SEM microphotographs obtained over the Ru catalysts of this study; EPR spectra obtained at room temperature (298 K) over 6Ru/SiO2, 6Ru/CePrO, and 6Ru/HEO catalysts; comparative raman spectra of Ru/CePrO and Ru/HEO catalysts; deconvolution of Ru 3p core-level spectra of Ru/CePrO and Ru/HEO; deconvolution of O 1s core-level spectra of Ru/CePrO, Ru/CeLaO and Ru/HEO catalysts; areas of the O 1s deconvoluted peaks; Ru dispersion and particle size as obtained from the H2 chemisorption studies; and FFT analysis of the Ru/HEO catalyst (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.