Next Article in Journal
Preparation and Electrocatalytic Activity of Bimetallic Ni-Cu Micro- and Nanoparticles
Previous Article in Journal
Neodymium-Doped Zinc Oxide Nanoparticles Catalytic Cathode for Enhanced Efficiency of Microbial Desalination Cells
Previous Article in Special Issue
Engineering Heterostructures of Layered Double Hydroxides and Metal Nanoparticles for Plasmon-Enhanced Catalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthesis and Specific Properties of the Ceria and Ceria-Zirconia Nanocrystals and Their Aggregates Showing Outstanding Catalytic Activity in Redox Reactions—A Review

Department of Chemical Technology, Faculty of Chemistry, Jagiellonian University, Gronostajowa 2, 30-387 Krakow, Poland
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(8), 1165; https://doi.org/10.3390/catal13081165
Submission received: 6 July 2023 / Revised: 25 July 2023 / Accepted: 27 July 2023 / Published: 29 July 2023

Abstract

:
Non-stoichiometric CeO2−y, especially in the form of nanocrystal aggregates, exhibits exceptional catalytic activity in redox reactions. It significantly improves the activity of transition metals and their oxides dispersed on/or in it, also acting as an oxygen buffer. Particularly, active oxygen species (O2n−, O) are generated at the M/CeO2−y nanoparticle interface, as well as in the surface layer of their solid-state solutions MxCe1−xO2−y. The crystal structure of CeO2, ZrO2 and (Ce, Zr)O2 and its defects are discussed in connection with the resulting specific catalytic activity. All the methods (simple precipitation and co-precipitation from mother liquors, sol–gel methods, precipitation from nanoemulsions, hydrothermal and solvothermal techniques, combustion and flame spray pyrolysis, precipitation using molecular and solid-state matrices, 3D printing and mechanochemical methods) used for the synthesis of these nanomaterials are comprehensively reviewed, describing the rules of individual procedures and preparation details. Methods of deposition of metal catalysts and their oxides on CeO2 nanoparticles, such as impregnation, washcoating and precipitation deposition, were also discussed. This review contains more than 160 references to representative papers wherein the reader can find further details on individual syntheses of effective ceria-based catalysts for redox reactions.

1. Introduction

A desired property of most catalyst carriers is ability to form high-surface-area materials with appropriate porous structure, e.g., mesoporous or hierarchical, on which high-surface-area catalytic precursors are dispersed. During subsequent chemical and temperature treatments, these precursors are transformed into embedded nanoparticles with active sites for a given catalytic reaction. During these treatments and subsequent treatments as a catalyst, these nanoparticles should not be sintered, thereby losing active sites. In addition, the porous structure of catalytic carriers should provide reactants with easy access to the active sites on the inner pore walls, as well as easy diffusion of the reaction products back to the gas phase.
The nanocrystal aggregates of CeO2−y, or solid-state solutions Ce1−zZrzO2−y, or mixed oxides CeO2-ZrO2 obtained in the developed synthesis methods show mesoporous [1,2,3,4,5] and hierarchical [6,7,8] structures satisfactorily meeting these requirements. Sometimes, mesoporous ceria-based catalysts MxCe1−xO2−y are obtained in a single step, for example, using reverse nanoemulsion methods [9,10,11,12,13,14,15].
Moreover, these nanocrystals exhibit catalytic activity in redox reactions and even in cases of noble metals or transition metal oxides dispersed on these ceria-based aggregates, lattice oxygen and oxygen vacancies of the ceria lattice can participate in one or more steps of the catalytic circle proceeding with these complex catalysts [3,14,16,17,18]. These active catalytic supports can also help in gasification of carbon deposits poisoning oxidation or reforming catalysts (especially dry reforming) of organic compounds [18,19,20,21]. CeO2−x additives also help to reactivate diesel filters by oxidizing carbonaceous deposits that build up during engine operation [22]. Such catalytic activity is expressed in the ability to donate oxygen under reducing conditions and take it back under oxidizing conditions [3,23,24,25]. The limited changes in the stoichiometry caused by this participation usually do not cause changes in the essential fluorite-type structure of CeO2 and ZrzCe1−zO2−y until z = 0.2, although as z increases, more and more deformation occurs in CeO6 octahedrons [26]. However, the progress of the reduction can change nanocrystal face exposures [27,28,29].
Another important feature of CeO2 is the ability to incorporate various active transition metal ions, creating solid solutions MxCe1−xO2−y within quite wide limits of x [1,3,4,5,11,24,30,31]. Such systems are characterized by both the variability of the M and Ce oxidation number and the concentration of anion vacancies and even the possibility of forming particularly active oxygen species, peroxides O22− and superoxides O2 [3,14,17,32,33]. Experimental evidence for the involvement of these active oxygen species using near-ambient-pressure XPS (NAP-XPS) has been compiled in a monograph written by Llorca’s research group [34].
Therefore, the characterization of ceria-based catalysts usually includes the measurement of the oxygen storage capacity (OSC) [3,5,14,15,17]. Exceptionally high OSC values allow ceria-based catalysts to maintain the stationarity of catalyzed reactions during their operation; they act as an oxygen buffer. This is supplemented by exceptionally high oxygen mobility owing to the fast oxygen transport pathways in defected CeO2−y, Ce1−zZrzO2−y, MxCe1−xO2−y and MxCe1−x−zZrzO2−y crystal lattices, so they are used as a solid electrolyte in high-temperature fuel cells [35,36,37]. These pathways are most probably used in the generally accepted Mars–van Krevellen mechanism of catalytic redox reactions over ceria-based catalysts [3,17,38,39,40]. The fast oxygen transport pathways enable using oxygen and oxygen vacancies not only from the superficial layer but also from the deeper layers of the catalysts [41]. This explains the exceptionally high values of OCS in ceria-based catalysts.
The OSC is most often measured by temperature-programmed reduction and oxidation (TPR and TPO) or by isothermal titration with probe molecules (H2, CO, O2, N2O) [42,43,44,45,46]. Ceria-based catalysts with a particularly high OSC are used in processes characterized by rapid and large changes in the redox potential of reaction mixtures, primarily in car catalysts. These catalysts typically contain precious metals (Pt, Pd, Rh) embedded on a ceria-zirconia carrier that covers the surfaces of a metal mesh or cordierite ceramic support to provide good gas flow. An extensive review of this type of catalyst has been presented by Reddy’s research group [47]. High OSC is also helpful in inhibiting deactivation of catalysts owing to deposition of inactive carbon deposits [20,48,49] or, in cases of excess depth, reduction of metal ions during catalyst activation or operation [50,51].
The listed features of these ceria-based catalysts mean that they are very often used in green chemistry processes, limiting the harmful anthropogenic impact on the environment. They are commonly used to neutralize exhaust gases from vehicles with combustion engines or from power plants and combined heat and power plants. In many processes of chemical technology, harmful by-products are generated, e.g., volatile organic compounds, which must be incinerated. Here, the crystal and porous structure of ceria-based nanomaterials, methods of their preparation and their specific properties are presented.
Although there are reviews on ceria-based catalysts, there seems to be no one that covers a review of all the methods used to obtain these cerium oxide-based nanomaterials. The aim of our review is not only to fill this gap, but also to link particular methods of synthesis of nanocrystals and their aggregates with their specific crystal structure, their defects and catalytic activity in redox reactions. This review contains many references to original papers, allowing the reader to become acquainted with the details of particular methods of synthesizing ceria-based nanocatalysts and their application in exemplary reactions.

2. Concise Description of the Crystal Structure of CeO2−y, ZrO2−y and ZrzCe1−zO2−y

Crystal structures, their defects, phase transitions and the phase diagrams of CeO2−y, ZrO2−y and ZrxCe1−xO2−y systems were reviewed in detail by Yashima [26]. In the phase diagram of CeO2−y within 0 < y < 0.5, at least six phases were recognized which differ in molar ratio O:Ce and in space groups from fluorite-type to triclinic. The closer the ratio O:Ce is to 2:1, the more likely it is there is only the fluorite phase. The cubic fluorite structure of CeO2 belonging to the space group Fm3m is formed of CeO6 octahedrons and OCe4 tetrahedrons. Normally, ceria preparations show some kind of oxygen deficit, i.e., oxygen vacancies, which is compensated by partial reduction of Ce4+ to Ce3+. The last ion has a radius about 15% larger, which corresponds to an approximately 50% larger volume than Ce3+. An increase in lattice parameters with the reduction of ceria can be observed even in case of solid solutions MxCe1−xO2−y despite the formation of oxygen vacancies at the same time [11,13]. The most common impurity of the CeO2 fluorite phase is the Ce11O20 phase belonging to the P1 group [52]. This phase is characterized by a 30% formal reduction of Ce4+ to Ce3+ and 15% concentration of oxygen vacancies.
An increase in temperature intensifies the vibrations of Ce-O bonds anisotropically in the {111} direction, as was shown in calculated electron density maps [53]. These vibrations cause deformation of the polyhedrons, which in combination with the already existing anion vacancies creates two pathways of easy diffusion of oxide anions in the {111} direction and the z axis.
Zirconia (IV) oxide shows three clear phase transitions but at high temperatures that do not occur in catalytic processes. According to Yashima and Mitsuhashi [54], the first-order martensitic transformation from the monoclinic to the tetragonal phase only occurs at 1400 K and is characterized by high hysteresis. The monoclinic phase of ZrO2 shows baddeleyite structure (P21/c), which can be described as such a strongly deformed fluorite structure for which the coordination number of zirconium (CNZr) is equal to 7. The tetragonal zirconia phase (P42/mcm) can be imagined as a little less deformed fluorite structure with CNZr = 8. At a very high temperature (2642 K), this tetragonal phase is transformed into the real cubic fluorite structure with CNZr = 6.
A common addition to CeO2 is ZrO2; this oxide itself is thermally much more stable which improves the thermal stability of Mx(ZrCe)1−xO2−y materials in relation to MxCe1−xO2−y. This is owing to the fact that M/(ZrCe)1−xO2−y and Mx(ZrCe)1−xO2−y catalysts can be applied in high-temperature processes, e.g., in the exhaust systems of vehicles, which is discussed in detail in a review published by Reddy’s group [47]. Numerous research papers by Yashima’s team, cited in [54], allow one to accurately determine the phase diagram of ZrzCe1−zO2 within a wide range of z. Within 1≤ z ≤ 0.92, there is a single monoclinic phase of ZrO2. Within the next narrow range of 0.92 ≤ z ≤ 0.88, an admixture of tetragonal ZrO2 occurs, for which the t phase above z = 0.88 is present only up to z = 0.80. Above this value of z, namely within 0.80 ≤ z ≤ 0.40, only the next tetragonal phase t′ exists. Within 0.40 ≤ z ≤ 0.35, the concentration of the third tetragonal phase t″ gradually increases. With further increases in Ce concentration, only one tetragonal phase t″ exists up to z = 0.10. Below this value of z, only the ceria cubic phase exists in the system. In a later study, Yashima [55], using high-angle X-ray diffraction, showed that the size of the crystalline ZrzCe1−zO2 system can change the range of existence of these t phases.
Although the mobility of anions in zirconia lattices is much lower than in the ceria lattice, even a small addition of Zr to the ceria lattice improves the anion mobility [56]. This is due to the deformation of the cubic structure of ceria in the direction of pseudo-cubic or pseudo-tetragonal structure where z > 0.1. This involves a slight shift of the O ions towards their position in the tetragonal lattice while the Ce and Zr ions remain in the positions of the cubic lattice. This deformation, impossible to notice in standard powder XRD, was observed in Raman spectroscopy and EXAFS by Yashima and Morimoto [57] and confirmed later using the same spectroscopic analyses by Vlaic for Rh-ZrzCe1−zO2 catalysts [58].
The DFT calculation performed on the super cell Ce8O16 of the fluorite structure showed that gradual atom-by-atom removal of oxygen causes lengthening of some—and shortening of other—Ce-O bonds, thus creating pathways of the fast diffusion of oxygen [59]. A similar calculation on the Zr8O16 super cell of tetragonal structure showed an analogue formation of the fast diffusion pathways with advance of gradual removal of oxygen atoms [53]. The experimental measurements of oxygen diffusion as a function of the ZrzCe1−zO2 system’s composition showed the highest oxygen mobility for equimolar composition Zr0.5Ce0.5O2 [56] which was in accordance with prediction coming from the calculated maximum of the molecular entropy in the ZrzCe1−zO2 system [60]. The results of the neutron diffraction measurements showed a shift of oxygen atoms in the {111} direction which creates pathways for the fast oxygen diffusion in {100} and {110} directions [61].
Surface reduction of ceria nanocrystals may change termination faces between (001), (110) and (111) [27,29]. Differences in reducibility of individual crystal faces of CeO2, (001), (110) and (111) have been investigated by Yuan and Haile and related to differences in the surface concentration of Ce3+ [29]. Such differences in the reduction of different faces were also observed during etching of ceria nanocrystals in liquid cell TEM [62]. Surface reduction may take place during thermal aging used to activate the ceria-based catalysts, causing structural changes at least in their surface layers [63]. Re-faceting of the exposed surface of ceria-based nanocatalysts may also occur during catalyzed reactions, especially of the redox type [27,28]. A mechanism for such surface re-faceting has been proposed by Yang et al. [64]. As was shown by Montini [65] during the deep reduction of the equimolar oxide system ZrO2-CeO2, a new phase Ce2Zr2O7+x appears of pyrochlore structure (Na,Ca)2Nb2O5(OH,F). The Ce2Zr2O7+x phase may be presented as a pseudo-cubic one where cations are located at corners and centers of the cube’s faces and seven oxygen ions are stochastic located at the tetrahedrons positions. This phase undergoes reversible redox cycles within a wide range of x.
In search of fast conductors of oxygen ions, the cubic cerium oxide systems containing admixtures of the rare metal R were frequently investigated. It was found that, in all cases, conductivity occurs through the hopping of O ions to adjacent anion vacancies and the highest conductivity occurred for composition R0.2Ce0.8O1.9 [56]. Among rare earth elements, the highest conductivity was shown for addition of Gd and Sm. However, in case of Y admixtures, the conductivity maximum occurred for a different composition, i.e., Y0.07Ce0.93O2−y [66]. Studies of electrical conductivity and impedance spectroscopy showed that the conductivity of CuxCe1−xO2−y solid solutions increased with Cu doping and is both electronic and ionic [67]. It has been shown that the catalytic activity in the total combustion of volatile organic compounds (VOCs) depends mainly on oxygen ions’ conductivity and the ion-hopping mechanisms of this conductivity have been confirmed.

3. Synthesis of the Ceria-Based Nanoparticles

Nanomaterials are usually defined as materials composed of nanoparticles with at least one dimension d < 100 nm. They are characterized by a significantly increased number of atoms located on the surface in relation to those located in the bulk of materials. In the case of nanocrystalline nanomaterials, less stable faces and atoms or ions located on the edges and vertices of intersecting faces are exposed in greater numbers. In these sites, atoms or ions appear that are significantly unsaturated in valence and, thus, predestined to be active sites of catalytic reactions. Xie et al. [68] showed this for the intersection of nanocrystal facets (111) and (002) of CuO/CeO2 nanocrystals obtained by the hydrothermal method (described later). The highest concentration of oxygen vacancies was also close to these intersections. Surface and volumetric structural and chemical defects should be added here. In the case of CeO2-based nanocrystals, oxygen vacancies and reduced Ce3+ ions play an important role [14,69]. This role may also be played by surface defects; in the case of CeO2−y, they are oxygen vacancies and reduced Ce ions. Therefore, partially reduced nanometric ceria mesoporous systems with a large specific surface area are catalytically valuable.
Methods of obtaining such nanomaterials can be divided generally into chemical and mechanochemical methods. Chemical methods are usually performed in two stages. During the first stage, a nanometric carrier CeO2−y or ZrxCe1−xO2−y is obtained. Such carriers already have their own catalytic activity. In the second stage, precursors of noble metals or transition metal oxides are deposited on the ceria-based carrier obtained previously. Mechanochemical methods are usually single-step, although it is not uncommon for additional fragmentation of nanomaterial agglomerates to be necessary.
The simplest of the chemical methods to obtain CeO2−y or ZrxCe1−xO2−y nanoparticles is precipitation or, respectively, co-precipitation of hydroxide and/or hydroxyoxide aggregates by a carefully controlled addition of a pH-raising agent to a solution of ceria nitrates or chlorides, usually. After some time of maturation of these aggregates, they are separated by centrifugation from the saturated mother liquors and washed, dried and calcined, to produce ceria carriers. In the case of co-precipitation, the precipitates of the metal ions used must have comparable solubility products. By manipulating the parameters of precipitation and subsequent operations, materials with the desired nanometric particle size can be achieved. They may, however, turn out to be phase-inhomogeneous. An example is the flower-shaped CuzZrxCe1−x−zO2−y catalyst which contains a cerium-enriched cubic phase and a zirconium-enriched tetragonal phase [70]. Nevertheless, these biphase catalysts, also called mixed oxides, were satisfactorily active in the preferential oxidation of CO.
Optimization of cerium concentration, anion type, pH values, process temperature and time can lead to obtaining nanoparticles of the desired shape. The driving forces here are the structural anisotropy of the chemical compound and the chemical potential of the mother liquor. For example, Mai [71] used the high asymmetry of Ce(III) hydroxide in the precipitate to produce CeO2 nanorods after drying and calcining. It was necessary to add concentrated (6–9 M) NaOH to the Ce(NO3)3·6H2O solution to accelerate the precipitation of Ce(OH)3 nuclei. Less concentrated solutions and increased temperature finally resulted in CeO2 nanocubes. However, Wu et al. [72] and Wang et al. [73] reported that chloride ions interacting on the (111) and (100) faces of growing crystals cause increases in the proportion of nanowires and nanorods (1D growth preferred), while NO3 anions favor the formation of nanocubes.
It would seem that these various statements can be explained by the results of the study by Liu et al. [74]. These authors observed that starting with cerium(III) nitrate or chloride in both cases yielded nanorods growing in the {111} direction. The differences are manifested in the exposure of the faces: (100) and (110) for nitrates and (111) and (100) for chlorides. However, this is not a certain rule because, by doing the same, Agarwal et al. [75] found only sidewall exposures (111). Therefore, there are reports of various shapes of terminal faces of nanorods, most often hexagonal, pentagonal and rectangular. The influence of the anions used in the synthesis on the shape of the obtained crystallites and their aggregates is not unambiguous, because the other factors affecting it are mutually intertwined. Therefore, there are still publications focused on this problem [76,77,78,79].
Taking into account the crystal structure of CeO2, the most stable structure of its nanocrystals should be in the shape of octahedron. However, it is often not thermodynamics that determines the shape of growing nanocrystals, but the kinetics of this growth. Then, octahedral crystals with truncated corners (tetradekaedrons) are formed as well as other less symmetrical crystal shapes. Therefore, apart from faces belonging to the (111) family, facets belonging to other face families are also displayed.
Crystal growth can be controlled by the addition of various inorganic and organic substances. For example, Zhang et al. [80] showed that a small addition of decanoic acid, which preferentially binds to (100) facets, inhibited the growth of nanocrystals in the {100} direction and accelerated in the {111} direction, resulting in the formation of nanocubes. However, larger additions of this acid also inhibit the growth in the {111} direction, thus causing the formation of smaller nanoparticles in the shape of octahedrons with truncated corners. It should be added that the frequent structural defects in nanorods make it difficult to unambiguously interpret their exact morphology. Recently, it was possible to control the ratio of exposed faces (111)/(110) in nanometric cup-shaped CeO2 particles [81].
It is also possible to force the desired shape of ceria nanoparticles, e.g., in the form of nano-needles, by adding Al3+ ions introduced into the CeO2 network [82]. In turn, the introduction of Eu ions promoted the growth of EuxCe1−xO2 nano-cups (x < 0.4), while improving the redox activity [83]. In the range of higher Eu concentrations (x > 0.4), the formation of ordered defect structures was observed. It has also been shown that the presence of water on the surface of CeO2 nanoparticles affects the evolution of their shape [84]. Xie et al. [68] prepared a series of CuO/CeO2 nanocatalysts with octahedron, rod, cube, sphere and spindle morphologies. The highest activity in CO oxidation was found for intersection of nanocrystals planes and, among them, the highest concentration of oxygen vacancies showed intersection of (111) and (002) facets.
Ceria as a semiconductor can be doped by aliovalent cations and anions in relation to the Ce4+ and O2− ions causing redox changes of these ions, generating point, geometric and chemical defects, changing the electrical conductivity and also creating active sites for catalytic reactions [85]. In case of the doped CeO2-based catalysts, anion vacancies, redox pair Ce 4+/Ce3+ and transition metal single-ion sites Mn+/(n−1)+ are generated [12]. All of them, depending on the reaction, may take place in the catalytic circle. Cop et al. [15] tried doping mesoporous CeO2 with both isovalent Zr and aliovalent Gd, Pr and Tb ions to optimize catalytic activity in the HCl oxidation reaction. There are reports on changing the shape of growing crystals of CeO2 caused by different doping methods with metal ions [77].
In addition, changes in the shape and exposure of the walls may arise at the calcination stage. Ta et al. [86] observed a change in the exposure of walls (110) and (100) to (111) after increasing the calcination temperature from 400 to 700 °C. Bezkrovnyi et al. [87] observed temperature-activated reconstruction of CeO2 nanocubes into a zig-zag structure of (111) nanofacets which increased the catalytic activity in CO oxidation.
Also, the reactants of catalyzed reactions may cause structural changes in the CeO2 lattice as was showed by Yang et al. [88]. These authors, using CO adsorption (IRRAS) and HRTEM on single crystals with exposed (111) and (110) faces, observed the reconstruction of a large part of the (110) faces into (111) faces. Matsukawa et al. [26] noticed a structural transition of CeO2 with increasing temperature in the presence of hydrogen related to partial bulk reduction of CeO2. Wang et al. [63] studied the activating effect of elevated annealing temperature on the formation of active sites in Cu/CeO2 catalysts. Yuan and Haile [29] noticed a strong dependence of the surface reduction progress on the type of exposed walls, i.e., (001) and (110) or (111). Polo-Garzon et al. [28] described and discussed this surface reconstruction for various metal oxides and its influences on their catalytic activity. Recently, Yang et al. [64] proposed a mechanism of surface re-faceting in cubic ceria nanocrystals.
Llorca’s group observed the reorganization of bimetallic nanoparticles caused by the catalytic reaction taking place on them [89]. These were RhPd particles deposited on ceria nanocubes and nanorods on which steam reforming of ethanol was carried out. The dynamics of this reorganization depended on the type of exposed ceria face. Li et al. [90] investigated the effect of the shape of CeO2 nanoparticles on the surface diffusion of Ru deposited on them. Liu et al. [91] observed that 2 nm NiO nanoparticles deposited on the (111) face of CeO2 at 500 K and in contact with CH4 migrate and form solid solutions NixCe1−x+O2−y. Liu et al. [92] noticed that the formation of alloys between metallic Ni and Cu nanoparticles deposited on a CeO2 support clearly depends on the morphology of the support used. Wang et al. [93] investigated the structural stability of various CeO2 structures during catalytic steam reforming of ethanol on Ir/CeO2 nanocatalysts. Differences in sintering of Au nanoparticles on flat faces (100) and (111) and on zig-zag nanofacetted structures on (111) faces were also observed [94]. These differences were related to differences in the preferential oxidation of CO. This is related to the direct observation by Sung et al. [62], namely an increased redox sensitivity of nanofacets on nanocrystalline ceria revealed during etching in liquid cell transmission electron microscopy. During CO oxidation over doped ceria nanocatalysts, Sartoretti et al. [95] observed evolution of defects in these nanocatalysts using in situ Raman spectroscopy.
The use of urea as a precipitating agent, which, when added to a solution of cerium-zirconium salts and hydrolyzed at a temperature close to the boiling point of the solution, leads to the precipitation of mixed hydroxyoxide with gelling additives, was described in US patents [96,97]. The next steps were thermal maturation of the gel and removal of water from the gel with a solvent with a low surface tension, e.g., dry 2-isopropanol, after which the material was dried and calcined in a flow of dry air or a suitable mixture containing Ar, O2 and CO2. Finally, mixed oxides of x(ZrO2)·(1 − x)CeO2 nanocrystals with a size of 4–9 nm and their aggregates of 100–1000 nm were produced. Other descriptions of such carriers’ precipitation with urea can be found, for example, in [98,99]. Wolski, et al. [100] compared the catalytic activity of Au-CeO2 nanocomposites co-precipitated rapidly from mixed nitrates by the addition of NaOH with catalysts co-precipitated slowly after addition of hydrolyzed urea or hexamethylenetetramine. The rapidly precipitated nanoparticles were finer, more homogeneous in size and more active in low-temperature oxidation of benzyl alcohol. Mileva et al. [101] obtained a mesoporous Ti-Ce oxide nanomaterial using co-precipitation with urea. It was found that the Ti:Ce ratio determined catalytic activity in the methanol decomposition and ethyl acetate oxidation.
Nanometric CeO2−y and ZrxCe1−xO2−y mixed oxides with a specific surface area of up to about 130 m2/g are also obtained by the sol–gel method, i.e., modified Pechini method [102], originally designed for obtaining ceramic dielectrics. It starts with the preparation of a colloidal hydroxy-alkoxy sol, in which metal ions are dispersed, forming a gel [103]. Colloidal products are usually extracted in ultracentrifuges and, as pointed out by Li et al. [104], the large differences in density produced by them determine which colloidal nanostructures are separated, and this determines what the products of further processing will be. Properly applied slow drying and annealing leads to the removal of alkoxyl and hydroxyl ions, converting the material into a nanometric oxide or mixed oxide. The parameters that must be controlled during the process are the pH and stoichiometry of the solution, the amount of water added during hydrolysis, gel time, temperature and effective solvent removal [105]. It is assumed that the structure of the obtained gels is amorphous and only the maturation stage leads to the formation of a crystalline structure.
The use of surfactants significantly increases the surface area of these materials, but it relates to the complication and extension of the synthesis time due to the need to wash out the organic compounds with a mixture of water and acetone. Terrible [106] obtained cerium-zirconium oxide nanomaterials with a surface area of up to 235 m2/g using cetyltrimethylammonium bromide as a surfactant. The final materials were composed of nanoparticles (4–6 nm) and were highly porous (0.66 cm3/g). Recently, Chavhan, et al. [107] prepared CeO2 nanoparticles via a modified hydrothermal method using urea and NaOH. A structure-directing surfactant agent cetyltrimethylammonium bromide (CTAB) was applied to force growth from spherical to irregular polygonal morphology. The initial 18–20 ceria nanocubes after hydrothermal reaction exposed highly reactive (100) faces. Addition of CTAB caused a reduction in the size of CeO2 to 8–10 nm with the (111) exposition of these more stable (111) faces; at the same time, concentrations of oxygen vacancies and Ce3+ were enhanced in the cubic fluorite structure. This CeO2 nanomaterial obtained using the prompt procedure showed increasing structural disorder in the following order: nanocubes < nanoplates < polygonal structures < nanospheres. On the surface of CeO2−y and ZrxCe1−xO2−y nanoparticles obtained after the completion of the first stage of catalyst synthesis, i.e., after their calcination, noble and transition metals are deposited in the second stage of this synthesis. The simplest method is impregnation of these carriers with a solution containing salts or complexes of the catalytically active metals mentioned. After drying and thermal and chemical treatment, they remain as metal or metal oxide film covering the surface of the ceria-based support. An example may be CuO/CeO2 catalysts obtained by the wetness incipient impregnation method using a solution of Cu(NO3)2 and nanoparticles of CeO2−y [67]. These catalysts were significantly less active in catalytic incineration of VOCs than nanocrystalline solid solutions CuxCe1−xO2−y with the same Cu:Ce atomic ratio.
By using more advanced methods of covering nanoparticles with nanocrystals of another oxide, i.e., by decorating these nanoparticles, a material with a different, or even higher, catalytic activity is obtained than when implementing the component oxides separately. In our laboratory, we obtained nanocatalysts CeO2/Mn3O4 and CeO2/Co3O4 with a molar ratio Ce:M = 1:99 using the deposition precipitation method described below [3,5]. Although the 10–15 nm nanoparticles of Mn3O4 and Co3O4 were very active, decorating their surface with 2–5 nm CeO2−y nanocrystals (which were much less active separately) resulted in a significant increase in the catalytic activity in the oxidation of VOCs. Another example, reported by Huang et al. [17], is the synthesis of nanocrystalline CeO2−y decorated with β-MnO2 nanorods. These catalysts were distinguished by an increased oxygen transfer capacity, which is crucial for the activity of oxidation reaction catalysts using the Mars–van Krevellen mechanism. Mori et al. [108] obtained CoOx-decorated CeO2 heterostructures and noticed the dependence of catalytic activity in diesel soot combustion on the morphology of these catalysts.
Similarly, Liu et al. [92] prepared multi-metallic catalysts on different shaped CeO2−y using an impregnation method. They revealed dependence of Ni-Cu alloying behavior on the support morphology. Only the nanopolyhedron ceria made it possible to obtain homogeneous nanoalloys owing to the equivalent interactions of Ni and Cu species with CeO2−y (111) facets exposed on the nanopolyhedrons of CeO2. Such catalysts showed outstanding catalytic performance in acetylene and hexyne hydrogenation, confirming the importance of the proper tailoring of the support morphology.
Other modifications of simple impregnation have also been used to improve the uniform dispersion of deposited nanoparticles. Before impregnation of the cerium supports with an ammonia solution of tetraammineplatinum(II) in 2-propanol, they were pre-treated for several hours with organic acids (maleic or citric) [96]. Similarly, using a modified impregnation method, a WO3-TiO2/ZrxCe1−xO2 catalyst was obtained and successfully used in the selective catalytic reduction of NOx with ammonia (NH3-SCR) process with mixtures imitating the composition of exhaust gases from diesel engines [109]. Di Sarli et al. [22] applied a wash-coated diesel particulate filter (DPF) with highly dispersed nanometric CeO2. This filter allowed for total removal of the soot covering the filter starting from 475 °C and remaining stable for multiple coating and soot removal cycles. These authors [49] also compared the activity of SiC DPFs pre-coated with cerium oxide and then wash-coated with Cu and Ag nitrates with filters wash-coated with a suspension of nanometric CeO2 in metal nitrate solution. By far, the best catalyst for soot removal turned out to be the Ag-promoted DPF obtained by the second procedure, similar to the deposition–precipitation method.
The deposition–precipitation method consists in preparing a suspension of fine particles of a high-surface-area carrier, e.g., CeO2−y, to which a solution containing catalytically active transition metal salts is added. Then, a precipitating agent, e.g., ammonia solution, is gradually added dropwise, causing precipitation and deposition of hydroxyoxides on the surface of the carrier. After centrifugation of the solids, washing, drying and calcination, the active ingredient is highly dispersed on the surface of the carrier. Examples are a series of cerium-zirconium catalysts deposited with an NH3 solution on aluminum, titanium and silicon oxides that have shown catalytic activity in CO oxidation consistent with their oxygen storage capacity (OSC) values [110]. The deposition–precipitation method was used by Ta et al. [86] for depositing Au nanoparticles on CeO2−y. The use of this method has been extended to other M/CeO2 catalysts, especially Pt, Pd and Ni, by Cargnello et al. [111].
In a similar way, catalysts with transition metals deposited on CeO2−y and ZrxCe1−xO2−y nanoparticles were also obtained, for example, a series of K, Co/ZrxCe1−xO2−y catalysts using the sol–gel method with subsequent deposition–precipitation of Co and K on the solid solutions ZrxCe1−xO2−y [112]. Increases in the OSC values, catalysts’ sintering resistance and reducibility as well as activity in the oxidation of diesel soot with increased Zr contents were reported. The K promotion of Co/ZrxCe1−xO2−y resulted in the highest catalytic activity which the authors related to greater availability of active oxygen species, both adsorbed and networked. Alkali promotion (K and Cs) was frequently used to improve catalytic activity in redox reactions, for example, K in diesel soot combustion on Ce0.65Zr0.35O2 monoliths [113] and Na in N2O decomposition on CuO-CeO2 mixed oxides [78].
In our laboratory, MnxCe1−xO2−y and CoxCe1−xO2− nanoparticles were deposited on micrometric α-Al2O3 particles using the modified reverse nanoemulsion one-step technique, which enables direct preparation of nanocrystalline solid solutions and mixed oxides, as is described in the next paragraph [3,5,67]. These deposited nanoparticles consisting of solid-solution nanocrystals decorated and retained their specific catalytic activity on the inert α-alumina support.
The reverse microemulsion-mediated technique for formation of nanosized particles was proposed by Zarur and Jing [9]. Using this method, obtained nanosized particles should be much more homogeneous than those obtained by the other methods. Wang et al. [30] conducted a comprehensive structural study (synchrotron-based XRD, X-ray absorption and Raman spectroscopies and density functional calculations) to determine the possibility of incorporating Cu ions into the CeO2 cubic network and to determine their location in this network. The Rietveld refinement of the high-Q X-ray diffraction data as well as density functional calculations indicated Cu substitution of Ce in the cubic ceria lattice coexisting with formation of O vacancies. This substitution caused strong structural distortions manifested by the pseudo-8 coordination number of Cu. Only four of the oxygen atoms remain in the immediate vicinity, and the other two are 0.04 and 0.11 nm away [30]. Such strong structural distortions certainly contribute to the increase in the lattice parameter of CuxCe1−xO2−y with an increase in x despite the simultaneous increase in the concentration of O vacancies observed by Molenda et al. [11].
In our laboratory, we have developed a one-stage precipitation method for obtaining ceria-based MxCe1−xO2−y nanoparticles with several transition metals to check the formation of single-phase solid solutions (SPhSSs) by mixing two suitable inverted nanoemulsions (w/o) [11,12]. This method eliminates the stage of preparing the nanoparticles of ceria carriers in the first stage, on which precursors of metals or their oxides would be deposited, usually by the deposition–precipitation technique, in the second stage. Both nanoemulsions were prepared using cyclohexane as the oil phase and Triton X-100 as surfactant with n-hexanol as co-surfactant. Solutions of Ce and transition metal salts in appropriate ratios were nanoemulsions in the first and a water solution of the precipitating agents in the second. All the tested agents, namely TPAH (tetrapropylammonium hydroxide), TEAH (tetra ethylammonium hydroxide), NaOH, NH3 and (NH4)2CO3, proved to be effective, with a slight preference toward tetraalkylammonium hydroxides. The precipitation reactions took place in a large number of nanoreactors (10 <d < 50 nm), which were individual nanodrops of the water phase. After separation, drying and optimized thermal treatment, based on results of the thermal analysis coupled with mass spectra analysis of the evolved gas [12], nanocrystals (2 < d < 4 nm) forming highly porous aggregates (10 < D < 20 nm) were obtained.
The formation of nanocrystals of SPSSs was confirmed. The transition metal ions in the cubic CeO2 structure were solved up to a limited x value ranging (depending on the applied M) from about 0.15 to 0.25 in the formula MxCe1−xO2−y. Above these limits, a spontaneous separation or/and segregation of the MOx phase took place. Our trials to obtain the inverted solid solutions failed, so the solid solutions of CeO2 in Cu, Mn and Co oxide phases using the same modified method failed. From about 1 at% of CeO2, the strongest line (111) of this oxide appeared among the dominated lines of CuO, Mn3O4 and Co3O4 [3,5,67].
In the case of CuO-CeO2, single-phase solid solutions of Cu2+ in the cubic CeO2 are formed as nanocrystals of CuxCe1−xO2−y for 0.12 < x < 0.15. At x = 0.15, the reflections at 2 Θ of 35.5° and 38.8° characteristic for the CuO phase were found in the XRD patterns [12,67]. The reference catalysts CuO-CeO2 obtained by the wetness incipient impregnation method using the same starting solutions and commercial nanoparticles of CeO2 provided biphase nanoparticles, although the same conditions of precipitation, maturation, drying and calcination were maintained. The CeO2 pores were clogged by the CuO phase covering CeO2 nanoparticles, causing a drastic reduction in the surface area and, thus, a drastic loss in the activity during total oxidation of methanol. The measured T100 values were equal to 225 and 370 °C, respectively [67].
The CoxCe1−xO2−y solid solutions obtained by the modified reverse nanoemulsion method were the only phase up to about x = 0.2, when separate Co3O4 crystallization began, resulting in an appearance of the most intensive diffraction line (311) of Co3O4 at 2 Θ = 37°. References [3,114], reported the separation of Co3O4 in the course of CoxCe1−xO2−y preparation at the atomic ratio Co:Ce = 2:3, though the strongest (311) Co3O4 line was noticeable in their published XRD results at lower Co contents, even at Co:Ce = 1:4, as it was in our case [3]. Taking into account both of these results, one may assume that, in case of CoxCe1−xO2−, the separation of the Co3O4 phase begins when, on average, in a crystallite containing five cubic unit cells of CeO2, in four of those, one Ce atom is replaced with a Co atom, in other words, when, on average, 4 Ce atoms out of the 20 Ce atoms contained in the 5 CeO2 unit cells are exchanged for Co. In case of Co-rich nanoparticle materials obtained by the same method, no solution of Ce in Co3O4 was present; only a nanocomposite of the CeO2−y nanocrystals and twice-bigger Co3O4 nanocrystals were present in the range 0.85 < x < 1.
Synthesis of the solid solutions of MnxCe1−xO2−y using the one-step modified nanoemulsion method, described above, allows one to obtain nanoparticles composed of this SPSS up to about x = 0.25 [5]. Using the method of imaging, it can be assumed that the extraction of the spinel phase begins when, on average, in each unit cell of cubic MnxCe1−xO2−y phase, there is one Mn atom instead of a Ce atom. For CuxCe1−xO2−y, separation of the CuO phase begins when, on average, in each unit CeO2 cell, one half of a Ce atom is exchanged for Cu and, in the case of CoxCeO2−y, it shall happen if three-quarters of a Ce atom is exchanged for Co. Therefore, the highest solubility of the transition metal in CeO2 phase, measured by the xmax values, clearly depends on the maximum valence, at least of these three metals in their oxides.
The size of the single-phase crystals of MnxCe1−xO2−y was from 2 to 5 nm, increasing with x. After exceeding this limit value, the final product becomes two-phase; the Mn3O4 phase appears additionally. Synthesis of Mn-rich nanoparticles resulted in biphasic products confirmed by electron microscopy, and even at 1% Ce content, the most intense line (111) of CeO2 appeared in XRD patterns. The Mn-rich product should be presented rather as Mn3O4-CeO2 mixed oxides and not by the formula MnxCe1−xO2−y suggesting formation of any solid-state solutions.
Hydrothermal and solvothermal syntheses are preferred for the production of larger amounts of nanometric material. Hydrothermal synthesis consists in long-term heating in a closed autoclave at temperature close to the boiling point of the water solution of metal salts with the addition of a precipitating agent. By appropriately manipulating the process parameters (concentrations, temperature, time), nanomaterials with specific shapes and catalytic properties (OSC, activity and selectivity) can be obtained. It is also possible to obtain the SPhSS of ZrxCe1−xO2−y starting from salt and NaOH solutions and an oxidizing agent (H2O2 or NaBrO3) using the hydrothermal method [115]. Papadopoulous et al. [116] proposed a novel hydrothermal synthesis for Cu-promoted ceria catalysts using citric acid as a chelating agent and NaOH to obtain atomically dispersed Cu ions, which was confirmed by EPR analysis. These catalysts were characterized by a clearly increased catalytic activity.
Hydrothermal synthesis also allows one to obtain mesoporous materials. Xiao et al. [117], starting from an aqueous solution of metal salts, glucose and acrylamide and ammonia as a precipitating agent, obtained flower-like mesoporous materials Ce0.9M0.1O2−y, where M = Y, La, Zr, Pr and Sn. The synthesis products were initially calcined at 600 °C in Ar and then at 400 °C in air. In order to obtain the highest catalytic activity in CO oxidation, they were activated by a reduction in the H2/Ar stream at 800 °C. Si et al. [118] obtained, using the hydrothermal method, homogeneous solid-state solutions of trivalent rare earths in CexZr1−xO2 (x = 0.4–0.6). They reported that the larger ions of rare metals are more stable in the ceria-zirconia nanomaterial lattices. The effect of ammonia concentration during hydrothermal synthesis on the structure and redox properties of ZrxCe1−xO2 solid solutions was determined by Zhang et al. [119]. Using non-equilibrium supercritical hydrothermal synthesis, Zhu et al. [120] obtained highly Cr-substituted CeO2−x nanoparticles. These materials were characterized by sufficiently high oxygen storage capacity, which enabled the upgrading of bitumens at low temperatures.
An interesting modification of the hydrothermal method is its supplementation by the precipitation deposition. ZrxCe1−xO2−y nanotubes with a length of several hundred nm were obtained in such a way. A ZrO2 nanosuspension in a water solution of cerium nitrate with NaOH as a precipitation agent was conditioned for hours in a closed autoclave [66]. These nanotubes were highly active in the conversion of ethanol to hydrogen, which the authors attribute to the dominant exposure of the most active crystal faces (100) and (110). Zr0.5Ce0.5O2−y nanoparticles with a size below 10 nm, obtained using the supercritical solvothermal method in polyethylene glycol, had an exceptionally high OSC value [121].
Sometimes, the method of deposited precipitation is supplemented by hydrothermal treatment of the product. This provided a zirconium cerium catalyst supported on SBA-15 for the oxidative dehydrogenation of ethylbenzene to styrene using CO2 [122]. Due to the ordered mesoporous structure of SBA-15, this synthesis could also be qualified as co-precipitation using a solid template. Xi et al. [123] prepared bimetallic catalysts NixCo(10−x)/SBA-15 by co-precipitation deposition on a solid template using urea as a precipitation agent. The highest activity, coke and sintering resistance was determined for the 4–5 nm metal particles in the dry reforming of methane. Various structural and physicochemical analyses showed embedding of Ni and Co nanoparticles into the channel of SBA-15 with the formation of alloys. The highest activity, stability and H2/CO molar ratio (close to 1) measured during 50 h of operation at a gas hourly space velocity of 72,000 mL/(gcat h) under 973 and 1073 K, with no coke deposition, was found for the Ni9Co1/SBA-15 catalyst.
The greatest opportunities in controlling the synthesis of CeO2−y and MxCe1−xO2−y nanomaterials are provided using appropriate molecular templates (soft templates) made of liquid crystals or surfactant molecules organizing into a specific template depending on the applied parameters of the solution state (T, ci, pH). In this colloidal template, nanoparticles of precursors of oxide systems with shapes forced by the template are nucleated and grow. They are then subjected to washing, drying and thermal treatment in a controlled atmosphere leading to the transformation of the precursors into oxide systems and stabilizing their structure in future catalytic action. Examples of such nanostructured catalysts are CuxCe1−xO2−y [124] and Pt/Zr0.3Y0.1Ce0.6O2−y [125]. In the latter referenced work, Pt nanoparticles were embedded in a colloidal crystal template of yttrium-doped ceria-zirconia which in further processing turned into a macroporous catalyst Pt/Zr0.3Y0.1Ce0.6O2−y highly effective in low-temperature methane combustion. Hollow nanospheres or nanotubes can be produced in this way [126,127,128]. Using appropriate templates, researchers were able to produce hollow nanospheres or nanotubes.
Solids or hard templates with a hierarchical porous structure, such as SBA-15, zeolites or polymers, characterized by large specific surfaces and porosity, are also used. In such a way, Rombi et al. [129] obtained highly active NixCe1−xO2−y catalysts for CO methanation. Furthermore, Li et al. [130], using a combination of two templates, namely the molecular triblock copolymer Pluronic F127 and solid-state polymethyl methacrylate micro-spheres, obtained nanocrystals of Ce0.6Zr0.4O2−y and Ce0.7Zr0.3O2−y. These materials had a three-dimensional ordered macroporous structure with hollow mesoporous (warm-like) channels in the walls. The sizes of these nanocrystals were in the range of 4–7 nm. Rood et al. [131] used graphene oxide as a solid template to obtain enhanced ceria nanoflakes effective in CO oxidation and dry reforming of methane.
The use of templates requires their removal after synthesis, which is often not a trivial task. For this reason, methods that do not require the use of surfactants and templates are still used [17,68]. This is because by maneuvering only the parameters of hydrothermal and solvothermal processes, it is possible to obtain ceria-based nanoparticles of various shapes, for example, nanotubes [132,133,134] or hollow nanocubes, using peracetic acid as an oxidant and without templates [135].
During the synthesis of nanometric systems, there are always stages of precursor heating and calcination, which, especially in mass synthesis, are associated with temperature and concentration gradients of reagents. Reducing the impact of these factors on the heterogeneity of the obtained products can be achieved, especially in aqueous and alcoholic environments, by microwave heating, which additionally shortens the operation time and reduces the likelihood of contamination of the product by heaters and furnace walls. These issues were the subject of a monograph published by Reddy’s group with an emphasis on automotive catalytic converters [136]. The use of microwave heating allowed researchers to obtain a higher concentration of zirconium in the cubic fluorite-type Zr0.4Ce0.6O2−y solid-solution phase [137]. In this solution, a much greater share of oxygen atoms shifted to tetragonal positions, manifested in the Raman spectrum, were associated with exceptionally high catalytic activity in CO oxidation compared to catalysts heated in classic furnaces. Similar benefits are provided by the use of ultrasonic treatment in the synthesis of nanomaterial precursors, which was generally presented by Bang and Suslick [138]. In this way, Alammar et al. [139], starting from ionic liquids, obtained nanoparticles of CeO2−y characterized by increased activity in CO oxidation.
The often encountered thermal activated restructuring of the CeO2 surface also plays an important role. Tinoco et al. reported that annealing, in oxidizing conditions (873 K), of the exposed (100) walls leads to their restructuring into a zig-zag series of (111) facets [140]. The Au nanoparticles (2–3 nm) were preferentially located in the valleys between these (111) zig-zag walls on the periphery of the CeO2 nanoparticles with a previously restructured surface. Cargnello et al. [111] applied this procedure to other metals as well, including Pt, Pd and Ni. Also, the operating conditions of the catalyst may contribute to a significant reorganization of the bimetallic particles deposited on the cerium support. This was observed by Soler et al. [89] during steam reforming of ethanol into Rh and Pd/CeO2.
Combustion is often used for synthesis of nanomaterials as a fast method and shortens the sequence of synthesis processes, although it allows one to obtain materials with a not especially high surface area. It uses an oxidant, e.g., in the form of metal nitrates, and a reducer, usually organic fuel. Heating this mixture ignites the system and produces nanoparticles and exhaust gases. Sometimes, this simple combustion method results in high-surface-area materials. For example, Chen et al. [141] obtained mesoporous ceria-zirconia solid solutions with a specific surface area of 90–208 m2/g using a modified combustion method. The authors used salt-assisted combustion with ethylene glycol as fuel. The urea nitrate combustion method for the preparation of CuO/CeO2/ZrO2/Al2O3 catalysts was compared by Baneshi et al. [142] with the same composed catalysts obtained using a homogeneous co-precipitation method. The second method provided smaller Cu crystallites and better surface dispersion. In consequence, the catalytic activity in methanol steam reforming was higher, though the first catalysts produced less CO—which is more convenient for applying such produced hydrogen in fuel cells.
Great hopes are associated with flame spray pyrolysis, which allows for direct one-step procurement of nanoparticles. This method reduces the impact on the environment, shortens the time and lowers the price of catalyst production on a larger scale. Kim and Laine [143] obtained nanometric catalysts (Ce0.7Zr0.3O2)x(Al2O3)1−x[(Ce0.7Zr0.3O2)@Al2O3] of the core–shell type using organometallic precursors dissolved in ethanol and, after being converted into an oxygen aerosol, these catalysts were blown into a flame of combusted methane at a temperature above 1500 °C. The aerosol stream was then quenched at a rate of 1000 °C/s, and the product nanoparticles underwent electrostatic precipitation in an unaggregated form with a size not exceeding 20 nm. Strobel and Pratsinis [144] investigated the influence of the chemical composition of the precursor solvent on the morphology, structure and properties of the obtained nanoparticles. Sun et al. [145] described the use of several metal–organic frameworks (Ce-MOFs) synthesized by pyrolysis in toluene incineration as a representative of VOCs. The authors showed that the catalytic performance of catalysts with the smallest particles, the largest specific surface area, the highest concentration of surface oxygen species and the easiest reproducibility is best. These statements are, as it were, a summary of the general knowledge about the activity of oxidation reaction catalysts. Woźniak et al. [8] synthesized, by oxidative thermolysis of Ce0.9REE0.1(HCOO)3 mixed formate, star-like Ce0.9REE0.1O1.95 mixed oxides. These hierarchically structured star-shaped particles were compared with microemulsion-derived, loosely arranged nanoparticles. The proposed catalysts had higher activity in the soot combustion process.
Mechanochemical synthesis is often used to obtain nanomaterials, especially in the large-scale production of ceria-based catalysts. It consists in long-term intensive grinding in mills (reactive milling) of macro-grain mixtures of starting chemical compounds. This grinding leads to fragmentation, accumulation of energy, creation of defects and chemical reaction with formation of new reactive nanomaterials, e.g., mixed oxides. The general principles and application of this technique have been described in many review articles such as [146,147,148,149].
Mechanochemical synthesis is generally carried out at ambient temperature, which reduces energy consumption, and without organic solvents. Therefore, it is a green technology that complies with the principles of sustainable development of chemical technology. Considering this need, IUPAC included mechanochemistry as one of the 10 most important innovative technologies implementing this goal in the 21st century [150]. Moreover, due to the simplicity of the processing parameters, the adaptation of mechanochemical synthesis from lab benchtops to commercial-scale production took less than 8 years [150]. In order to facilitate the transfer of laboratory test results to an industrial scale, an attempt was made to standardize synthesis reports on a laboratory scale [151].
It is likely that Trovalleri et al. [152] were the first to use ball milling for the synthesis of mixed nanosized ceria-zirconia nanoparticles Ce0.5Zr0.5O2−y. They found that prolonging the milling beyond 9 h did not result in any further improvement in the product’s reducibility, which is related to the OSC and redox activity of the ceria-based catalysts. Frequently, steel ball mills contaminate the obtained nanoparticles with Fe and W. To avoid this, Cutrufello et al. [153] used a ceramic zirconium ball mill to obtain uncontaminated ceria-zirconia nanocatalysts for dehydrogenation of 4-methylpentan-2-ol. A review on the mechanochemical synthesis of ZrO2 nanoparticles was presented by Dodd and McCormick [154].
Often, a liquid component is added to the mill. For example, Zheng et al. [155], starting from ZrOCl2 ·2H2O, Ce2(CO3)3 and H2C2O4 ·2H2O, added various surfactants to ground substrates. The largest specific surface areas (>80 m2/g) of ZrxCe1−xO2−y nanoparticles were obtained using neutral surfactants and these carriers were the best for the subsequent deposition–precipitation of Pd and Rh in the production of TWC car catalysts. Tsuzuki and McCormick [156] showed that the addition of a significant excess of NaCl as a diluent to the stoichiometric mixture of CeCl3 and NaOH eliminated the often occurring ignition of the ground reagents during milling, which occurs even in Ar, resulting in the formation of large (500 nm) aggregates of CeO2−y. Li et al. [157] added higher amounts of NaCl and eliminated ignition, obtaining nanoparticles of size 40–70 nm. Starting from (NH4)2Ce(NO3)6 instead of CeCl3 and also using NaCl as an inhibitor of agglomeration, Li et al. [158] obtained ceria nanoparticles as small as 2.5 nm. The way NaCl acts to obtain highly active nanoparticles of CeO2 was discussed by Shu et al. [159] [Shu, Y. et al., 2020]. Cheng, H., et al. [160] presented a low-cost and solvent-free large-scale mechanochemical nanocasting method for obtaining highly porous CeMnOx catalysts for low-temperature NH3-SCR of NOx. These catalysts show high specific area and their reach in chemisorbed species allows for 100% conversion of NOx at 150 °C.
An innovative method of obtaining CeO2-based catalysts has recently been presented by the Llorca group [161]. They obtained ceria aggregates such as woodpile with micrometric channels from ceria powder paste with nitric acid using a 3D printer. The materials obtained in this way, after applying Ni on them, were tested in the ammonia decomposition reaction. They turned out to be more active under the same reactor conditions than the conventional catalysts of cordierite honeycomb wash-coated with Ni/CeO2.
Although this review focuses on the methods of preparing ceria-based catalysts, selected examples of individual syntheses also include the results of using the obtained catalysts in specific reactions. Most often it was CO oxidation, because it is the most common test of general redox activity, next to TPR and TPO. Information on the reactions tested in the cited publications is summarized in Table 1.

4. Conclusions and Outlook

Non-stoichiometric CeO2−y nanocrystals and their aggregates seem to be necessary ingredients of redox catalysts’ operation according to the Mars–Van Krevelen mechanism. This is due to the high reversible OSC values obtained by maintaining the essential crystal structure during redox cycles. This makes it possible to act as an oxygen buffer during catalytic reactions. In addition, active oxygen species (O2n−, O) are generated on active oxygen vacancy–Ce3+ complexes. This is particularly important in low-temperature processes of complete oxidation of volatile organic compounds, protection against loss of catalyst activity as a result of deposition of carbon deposits blocking active sites and regeneration of catalysts. Due to its thermal stability, CeO2, and especially its ZrxCe1−xO2 solid-state solutions and mixed oxides CeO2-ZrO2, can act as a stable catalytic carrier and activity improver of the transition metal or its oxide. The crystal lattices of both of these oxides are relatively flexible, able to accept transition metal ions, ensuring their high dispersion and protecting them against sintering. It can be assumed that their use as active ingredients in many different new catalytic processes will spread in the future.
This review presents all the methods used to obtain ceria-based nanocrystalline catalytic systems, both one- and two-stage, starting from simple methods of precipitation and co-precipitation and ending with a 3D printing method. The methods of deposition of nanoparticles of active noble metals and transition metal oxides as well as the creation of solid-state solutions of transition metals in ceria lattices are also discussed. All these methods of catalyst synthesis are supported by representative examples of their application in specific reactions. We believe that this review will be helpful for the reader who wants to quickly get an overview of all the methods used for the synthesis of ceria-based nanocatalysts and to choose the right method for the development of a catalyst for a chosen reaction.

Author Contributions

Conceptualization, R.D.; writing—original draft preparation, R.D.; writing—review and editing, R.D., M.M. and L.C.; supervision, R.D.; project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

This is a review paper and, therefore, no new data were created.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Voskanyan, A.A.; Tsui, C.-K.J.; Chan, K.-Y. Durable ruthenium oxide/ceria catalyst with ultralarge mesopores for low-temperature CO oxidation. J. Catal. 2020, 382, 155–164. [Google Scholar] [CrossRef]
  2. Zhang, X.; Wang, D.; Jing, M.; Liu, J.; Zhao, Z.; Xu, G.; Song, W.; Wei, Y.; Sun, Y. Ordered Mesoporous CeO2-supported Ag as an Effective Catalyst for Carboxylative Coupling Reaction Using CO2. ChemCatChem 2019, 11, 2089–2098. [Google Scholar] [CrossRef]
  3. Dziembaj, R.; Chojnacka, A.; Piwowarska, Z.; Gajewska, M.; Świętosławski, M.; Górecka, S.; Molenda, M. Comparative study of Co-rich and Ce-rich oxide nanocatalysts (CoxCe1−xOy) for low-temperature total oxidation of methanol. Catal. Today 2019, 333, 196–207. [Google Scholar] [CrossRef]
  4. Akinnawo, C.A.; Bingwa, N.; Meijboom, R. Tailoring the surface properties of meso-CeO2 for selective oxidation of benzyl alcohol. Catal. Commun. 2020, 145, 106115. [Google Scholar] [CrossRef]
  5. Chojnacka, A.; Molenda, M.; Chmielarz, L.; Piwowarska, Z.; Gajewska, M.; Dudek, B.; Dziembaj, R. Ceria based novel nanocomposites catalysts MnxCe1-xO2/γ-Al2O3 for low-temperature combustion of methanol. Catal. Today 2015, 257, 104–110. [Google Scholar] [CrossRef]
  6. Hu, F.; Chen, J.; Peng, Y.; Song, H.; Li, K.; Li, J. Novel nanowire self-assembled hierarchical CeO2 microspheres for low temperature toluene catalytic combustion. Chem. Eng. J. 2018, 331, 425–434. [Google Scholar] [CrossRef]
  7. Moyer, K.; Conklin, D.R.; Mukarakate, C.; Vardon, D.R.; Nimlos, M.R.; Ciesielski, P.N. Hierarchically Structured CeO2 Catalyst Particles From Nanocellulose/Alginate Templates for Upgrading of Fast Pyrolysis Vapors. Front. Chem. 2019, 7, 730. [Google Scholar] [CrossRef] [PubMed]
  8. Woźniak, P.; Miśta, W.; Małecka, M.A. Function of various levels of hierarchical organization of porous Ce0.9REE0.1O1.95 mixed oxides in catalytic activity. Cryst. Eng. Commun. 2020, 22, 5914–5930. [Google Scholar] [CrossRef]
  9. Zarur, A.J.; Jing, J.Y. Reverse microemulsion synthesis of nanostructured complex oxides for catalytic combustion. Nature 2000, 403, 65–67. [Google Scholar] [CrossRef]
  10. Bumajdad, A.; Zaki, M.I.; Eastoe, J.; Pasupulety, L. Characterization of nano-cerias synthesized in microemulsions by N2 sorptiometry and electron microscopy. J. Colloid Interface Sci. 2006, 302, 501–508. [Google Scholar] [CrossRef]
  11. Molenda, M.; Dziembaj, R.; Chmielarz, L.; Drozdek, M.; Dudek, B.; Piwowarska, Z.; Rafalska-Łasocha, A. Synthesis and characterization of nanostructured Cu-doped CeO2 nanopowders prepared by the w/o microemulsion method. Pol. J. Environ. Stud. 2009, 18, 151–154. [Google Scholar]
  12. Dziembaj, R.; Molenda, M.; Chmielarz, L.; Drozdek, M.; Zaitz, M.; Dudek, B.; Rafalska–Łasocha, A.; Piwowarska, Z. Nanostructured Cu–Doped Ceria Obtained by Reverse Microemulsion Method as Catalysts for Incineration of Selected VOCs. Catal. Lett. 2010, 135, 68–75. [Google Scholar] [CrossRef]
  13. Zaitz, M.; Molenda, M.; Chmielarz, L.; Piwowarska, Z.; Dudek, B.; Walas, S.; Dziembaj, R. Influence of defect structure on catalytic activity of nanometric materials based on ceria-doped copper. Funct. Mater. Lett. 2011, 4, 165–169. [Google Scholar] [CrossRef]
  14. Wang, H.; Luo, S.; Zhang, M.; Liu, W.; Wu, X.; Liu, S. Roles of oxygen vacancy and O- in oxidation reactions over CeO2 and Ag/CeO2 nanorod model catalysts. J. Catal. 2018, 368, 365–378. [Google Scholar] [CrossRef]
  15. Cop, P.; Maile, R.; Sun, Y.; Khalid, O.; Djerdi, I.; Esch, P.; Heiles, S.; Over, H.; Smarsly, B.M. Impact of Aliovalent/Isovalent Ions (Gd, Zr, Pr, and Tb) on the Catalytic Stability of Mesoporous Ceria in the HCl Oxidation Reaction. ACS Appl. Nano Mater. 2020, 3, 7406–7419. [Google Scholar] [CrossRef]
  16. Torrente-Murciano, L.; Gilbank, A.; Puertolas, B.; Garcia, T.; Solsona, B.; Chadwick, D. Shape-dependency activity of nanostructured CeO2 in total oxidation of polycyclic aromatic hydrocarbons. Appl. Catal. B Environ. 2013, 132–133, 116–122. [Google Scholar] [CrossRef] [Green Version]
  17. Huang, X.; Beck, M.J. Size-Dependent Appearance of Intrinsic Oxq “Activated Oxygen” Molecules on Ceria Nanoparticles. Chem. Mater. 2015, 27, 5840–5844. [Google Scholar] [CrossRef]
  18. Mahamulkar, S.; Yin, K.; Agrawa, P.K.; Davis, R.J.; Jones, C.W.; Malek, A.; Shibata, H. Formation and Oxidation/Gasification of Carbonaceous Deposits: A Review. Ind. Eng. Chem. Res. 2016, 55, 9760–9818. [Google Scholar] [CrossRef]
  19. Jang, W.-J.; Shim, J.-O.; Kim, H.-M.; Yoo, S.-Y.; Roh, H.-S. A review on dry reforming of methane in aspect of catalytic properties. Catal. Today 2019, 324, 15–26. [Google Scholar] [CrossRef]
  20. Gao, X.; Wang, Z.; Ashok, J.; Kawi, S. A comprehensive review of anti-coking, anti-poisoning and anti-sintering catalysts for biomass tar reforming reaction. Chem. Eng. Sci. 2020, 7, 100065. [Google Scholar] [CrossRef]
  21. Alli, R.D.; Zhou, R.; Mohamedali, M.; Mahinpey, N. Effect of thermal treatment conditions on the stability of MOF-derived Ni/CeO2 catalyst for dry reforming of methane. Chem. Eng. J. 2023, 446, 143242. [Google Scholar] [CrossRef]
  22. Di Sarli, V.; Landi, G.; Lisi, L.; Di Benedetto, A. Ceria-coated diesel particulate filters for continuous regeneration. AIChE J. 2017, 63, 3442–3449. [Google Scholar] [CrossRef]
  23. Campbell, C.T.; Pedden, C.H.F. Oxygen Vacancies and Catalysis on Ceria Surfaces. Science 2005, 309, 713–714. [Google Scholar] [CrossRef]
  24. Dziembaj, R.; Molenda, M.; Zaitz, M.M.; Chmielarz, L.; Furczoń, K. Correlation of electrical properties of nanometric copper-doped ceria materials Ce1-x CuxO(2-δ) with their catalytic activity in incineration of VOCs. Solid State Ion. 2013, 251, 18–22. [Google Scholar] [CrossRef]
  25. Righi, G.; Righi, G.; Magri, R.; Selloni, A. Methane Activation on Metal-Doped (111) and (100) Ceria Surfaces with Charge-Compensating Oxygen Vacancies. J. Phys. Chem. C 2020, 124, 17578–17585. [Google Scholar] [CrossRef]
  26. Yashima, M. Crystal and electronic structures, structural disorder, phase transformation, and phase diagram of ceria-zirconia and ceria-based materials. In Catalysis by Ceria and Related Materials; Trovarelli, A., Fornansiero, P., Eds.; Imperial College Press: London, UK, 2013; Volume 2, pp. 1–45. [Google Scholar]
  27. Matsukawa, T.; Hoshikawa, A.; Ishigaki, T. Temperature-induced structural transition of ceria by bulk reduction under hydrogen atmosphere. CrystEngComm 2018, 20, 4359–4363. [Google Scholar] [CrossRef]
  28. Polo-Garzon, F.; Bao, Z.; Zhang, X.; Huang, W.; Wu, Z. Surface Reconstructions of Metal Oxides and the Consequences on Catalytic Chemistry. ACS Catal. 2019, 9, 5692–5707. [Google Scholar] [CrossRef]
  29. Yuan, W.; Haile, S.M. Insensitivity of the extent of surface reduction of ceria on termination: Comparison of (001), (110), and (111) faces. MRS Commun. 2020, 11, 1–6. [Google Scholar] [CrossRef]
  30. Wang, X.; Rodriguez, J.A.; Hanson, J.C.; Gamarra, D.; Martinez-Arias, A.; Fernandez-Garcia, M. Unusual physical and chemical properties of Cu in Ce1-xCuxO2 oxides. J. Phys. Chem. B 2005, 109, 19595–19603. [Google Scholar] [CrossRef]
  31. Lee, K.J.; Kim, Y.; Lee, J.H.; Cho, S.J.; Kwak, J.H.; Moon, H.R. Facile Synthesis and Characterization of Nanostructured Transition Metal/Ceria Solid Solutions (TMxCe1-xO2-δ, TM = Mn, Ni, Co, or Fe) for CO Oxidation. Chem. Mater. 2017, 29, 2874–2882. [Google Scholar] [CrossRef]
  32. Migani, A.; Neyman, K.M.; Illas, F.; Bromley, S.T. Greatly facilitated oxygen vacancy formation in ceria nanocrystallites. J. Chem. Phys. 2009, 131, 064701. [Google Scholar] [CrossRef] [PubMed]
  33. Soler, L.; Casanovas, A.; Urrich, A.; Angurell, I.; Llorca, J. CO oxidation and COPrOx over preformed Au nanoparticles supported over nanoshaped CeO2. Appl. Catal. B Environ. 2016, 197, 47–55. [Google Scholar] [CrossRef] [Green Version]
  34. Garcia, X.; Soler, L.; Divins, N.J.; Vendrell, X.; Serrano, I.; Lucentini, I.; Prat, J.; Solano, E.; Tallarida, M.; Escudero, C.; et al. Ceria-Based Catalysts Studied by Near Ambient Pressure X-ray Photoelectron Spectroscopy: A Review. Catalysts 2020, 10, 286. [Google Scholar] [CrossRef] [Green Version]
  35. Inaba, H.; Tagawa, H. Ceria-based solid state electrolytes. Solid State Ion. 1996, 83, 1–16. [Google Scholar] [CrossRef]
  36. Jaiswal, N.; Tanwar, K.; Suman, R.; Kumar, D.; Upadhyay, D.; Parkash, O. A brief review on ceria-based solid electrolytes for solid oxide fuel cells. J. Alloy Comp. 2019, 781, 984–1005. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Lin, J.; Singh, M.; Hu, E.; Jing, Z.; Raza, R.; Wang, F.; Wang, J.; Yang, F.; Zhu, B. Superionic conductivity in ceria-based heterostructure composites for low-temperature solid oxide fuel cells. Nano-Micro Lett. 2000, 12, 178. [Google Scholar] [CrossRef]
  38. Mars, P.; van Krevellen, D.W. Oxidations carried out by means of vanadium oxide catalysts. Chem. Eng. Sci. 1954, 3, 41–59. [Google Scholar] [CrossRef]
  39. Bielański, A.; Haber, J. Oxygen in catalysis on transition metal oxides. Catal. Rev. Sci. Eng. 1979, 19, 1–41. [Google Scholar] [CrossRef]
  40. Minico, S.; Scire, S.; Crisafulli, C.; Maggiore, R.; Galvagno, S. Catalytic combustion of volatile organic compounds on gold/iron oxide catalysts. Appl. Catal. B Environ. 2000, 28, 245–251. [Google Scholar] [CrossRef]
  41. Su, Z.; Yang, W.; Wang, C.; Xiong, S.; Cao, X.; Peng, Y.; Si, W.; Seng, Y.; Xue, M.; Li, J. Roles of Oxygen Vacancies in the Bulk and Surface of CeO2 for Toluene Catalytic Combustion. Environ. Sci. Technol. 2020, 54, 12684–12692. [Google Scholar] [CrossRef] [PubMed]
  42. Ray, S.P.; Nowick, A.S.; Cox, D.E. X-ray and neutron diffraction study of intermediate phases in nonstoichiometric cerium dioxide. J. Solid State Chem. 1975, 15, 344–351. [Google Scholar] [CrossRef]
  43. Knappe, P.; Eyring, L. Preparation and electron microscopy of intermediate phases in the interval Ce7O12-Ce11O20. J. Solid State Chem. 1985, 58, 312–314. [Google Scholar] [CrossRef]
  44. Dziembaj, R.; Łojewski, T. Evolution of oxide layer on the cobalt foil studied by temperature programmed reduction. React. Kinet. Catal. Lett. 1994, 52, 437–443. [Google Scholar] [CrossRef]
  45. Dziembaj, R.; Łojewska, J.; Łojewski, T. Application of metal oxidation and oxide reduction to determine superficial phases containing active centres in redox catalysts. Solid State Ion. 1999, 117, 87–93. [Google Scholar] [CrossRef]
  46. Łojewska, J.; Makowski, W.; Tyszewski, T.; Dziembaj, R. Active state of model cobalt foil catalyst studied by SEM, TPR/TPO, XPS and TG. Catal. Today 2001, 69, 409–418. [Google Scholar] [CrossRef]
  47. Devaiah, D.; Reddy, L.H.; Park, S.-E.; Reddy, B.M. Ceria-zirconia mixed oxides: Synthetic methods and applications. Catal. Rev. 2018, 60, 177–277. [Google Scholar] [CrossRef]
  48. Vogt, E.T.C.; Fu, D.; Weckhuysen, B.M. Carbon deposit analysis in catalyst deactivation, regeneration and rejuvenation. Angew. Chem. Int. Ed. 2023, 62, e202300319. [Google Scholar] [CrossRef]
  49. Di Sarli, V.; Landi, G.; Di Benedetto, A.; Lisi, L. Synergy Between Ceria and Metals (Ag or Cu) in Catalytic Diesel Particulate Filters: Effect of the Metal Content and of the Preparation Method on the Regeneration Performance. Top. Catal. 2021, 64, 256–269. [Google Scholar] [CrossRef]
  50. Dictor, R.A.; Bell, A.T. Fischer-Tropsch synthesis over reduced and unreduced iron oxide catalysts. J. Catal. 1986, 97, 121–136. [Google Scholar] [CrossRef]
  51. Lee, K.-M.; Kwon, G.; Hwang, S.; Boscoboinik, J.-A.; Kim, T. Investigation of the NO reduction by CO reaction over oxidized and reduced NiOx/CeO2 catalysts. Catal. Sci. Technol. 2021, 11, 7850–7865. [Google Scholar] [CrossRef]
  52. Zinkevich, M.; Djurovic, D.; Aldinger, F. Thermodynamic modelling of the cerium–oxygen system. Solid State Ion. 2006, 177, 989–1001. [Google Scholar] [CrossRef]
  53. Yashima, M.; Kobayashi, S.; Yasui, T. Crystal structure and the structural disorder of ceria from 40 to 1497 °C. Solid State Ion. 2006, 177, 211–215. [Google Scholar] [CrossRef]
  54. Yashima, M.; Mitsuhashi, T.; Takashina, H.; Kakihana, M.; Ikegami, T.; Yoshimura, M. Tetragonal-Monoclinic Phase Transition Enthalpy and Temperature of ZrO2-CeO2 Solid Solutions. J. Am. Ceram. Soc. 2005, 78, 2225–2228. [Google Scholar] [CrossRef]
  55. Yashima, M.; Sasaki, S.; Yamaguchi, Y.; Kakihana, M.; Yoshimura, M.; Mori, T. Internal distortion in ZrO2-CeO2 solid solutions: Neutron and high-resolution synchrotron x-ray diffraction study. Appl. Phys. Lett. 1998, 72, 182–184. [Google Scholar] [CrossRef]
  56. Yashima, M.; Takizawa, T. Atomic Displacement Parameters of Ceria Doped with Rare-Earth Oxide Ce0.8R0.2O1.9 (R = La, Nd, Sm, Gd, Y, and Yb) and Correlation with Oxide-Ion Conductivity. J. Phys. Chem. C 2010, 114, 2385–2392. [Google Scholar] [CrossRef]
  57. Yashima, M.; Morimoto, K.; Ishizawa, N.; Yoshimura, M. Diffusionless tetragonal-cubic transformation temperature in zirconia solid solutions. J. Am. Ceram. Soc. 1993, 76, 2865–2868. [Google Scholar] [CrossRef]
  58. Vlaic, G.; Fornasiero, P.; Geremia, S.; Kašpar, J.; Graziani, M. Relationship between the Zirconia-Promoted Reduction in the Rh-Loaded Ce0.5Zr0.5O2 Mixed Oxide and the Zr–O Local Structure. J. Catal. 1997, 168, 386–392. [Google Scholar] [CrossRef]
  59. Yashima, M. Crystal Structures of the Tetragonal Ceria−Zirconia Solid Solutions CexZr1−xO2 through First Principles Calculations (0 ≤ x ≤ 1). J. Phys. Chem. C 2009, 113, 12658–12662. [Google Scholar] [CrossRef]
  60. Yashima, M.; Wakita, T. Atomic displacement parameters and structural disorder of oxygen ions in the CexZr1−xO2 solid solutions (0.12 ≤ x ≤ 1.0): Possible factors of high catalytic activity of ceria-zirconia catalysts. Appl. Phys. Lett. 2009, 94, 171902. [Google Scholar] [CrossRef] [Green Version]
  61. Wakita, T.; Yashima, M. Structural disorder in the cubic Ce0.5Zr0.5O2 catalyst: A possible factor of the high catalytic activity. Appl. Phys. Lett. 2008, 92, 101921. [Google Scholar] [CrossRef]
  62. Sung, J.; Choi, B.K.; Kim, B.; Kim, B.H.; Kim, J.; Lee, D.; Kim, S.; Kang, K.; Hyeon, T.; Park, J. Redox-sensitive facet dependency in etching of ceria nanocrystals directly observed by liquid TEM. J. Am. Chem. Soc. 2019, 141, 18395–18399. [Google Scholar] [CrossRef]
  63. Wang, B.; Zhang, H.; Li, X.; Wang, W.; Zhang, L.; Li, Y.; Peng, Z.; Yang, F.; Liu, Z. Nature of Active Sites on Cu-CeO2 Catalysts Activated by High-Temperature Thermal Aging. ACS Catal. 2020, 10, 12385–12392. [Google Scholar] [CrossRef]
  64. Yang, C.; Capdevila-Cortada, M.; Dong, C.; Zhou, Y.; Wang, J.; Yu, X.; Nefedov, A.; Heißler, S.; López, N.; Shen, W.; et al. Surface Refaceting Mechanism on Cubic Ceria. J. Phys. Chem. Lett. 2020, 11, 7925–7931. [Google Scholar] [CrossRef] [PubMed]
  65. Montini, T.; Bañares, M.A.; Hickey, N.; Di Monte, R.; Fornasiero, P.; Kašpara, J.; Graziania, M. Promotion of reduction in Ce0.5Zr0.5O2: The pyrochlore structure as effect rather than cause? Phys. Chem. Chem. Phys. 2004, 6, 1–3. [Google Scholar] [CrossRef]
  66. Chen, Y.-C.; Chen, K.-B.; Lee, C.-S.; Lin, M.C. Direct Synthesis of Zr-Doped Ceria Nanotubes. J. Phys. Chem. C. 2009, 113, 5031–5034. [Google Scholar] [CrossRef] [Green Version]
  67. Dziembaj, R.; Molenda, M.; Chmielarz, L.; Zaitz, M.M.; Piwowarska, Z.; Rafalska-Łasocha, A. Optimization of Cu doped ceria nanoparticles as catalysts for low-temperature methanol and ethylene total oxidation. Catal. Today 2011, 169, 112–117. [Google Scholar] [CrossRef]
  68. Xie, Y.; Wu, J.; Jing, G.; Zhang, H.; Zeng, S.; Tian, X.; Zou, X.; Wen, J.; Su, H.; Zhong, C.-J.; et al. Structural origin of high catalytic activity for preferential CO oxidation over CuO/CeO2 nanocatalysts with different shapes. Appl. Catal. B Environ. 2018, 239, 665–676. [Google Scholar] [CrossRef]
  69. Chen, A.; Yu, X.; Zhou, Y.; Miao, S.; Li, Y.; Kuld, S.; Sehested, J.; Liu, J.; Aoki, T.; Hong, S.; et al. Structure of the catalytically active copper–ceria interfacial perimeter. Nat. Catal. 2019, 2, 334–341. [Google Scholar] [CrossRef]
  70. Moretti, E.; Storaro, L.; Talon, A.; Lenarda, M.; Riello, P.; Frattini, R.; De Yuso, M.D.V.M.; Jiménez-López, A.; Rodríguez-Castellón, E.; Ternero, F.; et al. Effect of Thermal Treatments on the Catalytic Behaviour in the CO Preferential Oxidation of a CuO-CeO2-ZrO2 Catalyst with a Flower-Like Morphology. Appl. Catal. B Environ. 2011, 102, 627–637. [Google Scholar] [CrossRef]
  71. Mai, H.X.; Sun, L.D.; Zhang, Y.W.; Si, R.; Feng, W.; Zhang, H.P.; Liu, H.C.; Yan, C.H.J. Shape-selective synthesis and oxygen storage behaviour of ceria nanopolyhedra, nanorods and nanocubes. J. Phys. Chem. B 2005, 109, 24380–24385. [Google Scholar] [CrossRef]
  72. Wu, Q.; Zhang, F.; Xiao, P.; Tao, H.; Wang, X.; Hu, Z.; Lu, Y. Great Influence of Anions for Controllable Synthesis of CeO2 Nanostructures: From Nanorods to Nanocubes. J. Phys. Chem. C 2008, 112, 17076–17080. [Google Scholar] [CrossRef]
  73. Wang, W.; Howe, J.Y.; Li, Y.A.; Qiu, X.F.; Joy, D.C.; Paranthaman, M.P.; Doktycz, M.J.; Gu, B.H. A surfactant and template-free route for synthesizing ceria nanocrystals with tunable morphologies. J. Mater. Chem. 2010, 20, 7776–7781. [Google Scholar] [CrossRef]
  74. Liu, X.W.; Zhou, K.B.; Wang, L.; Wang, B.Y.; Li, Y.D. Oxygen vacancy clusters promoting reducibility and activity of ceria nanorods. J. Am. Chem. Soc. 2009, 131, 3140–3141. [Google Scholar] [CrossRef]
  75. Agarwal, S.; Lefferts, L.; Mojet, B.L.; Ligthart, D.A.J.M.; Hensen, E.J.M.; Mitchell, D.R.G.; Erasmus, W.J.; Anderson, B.G.; Ezra, J.O.; Neethling, J.H.; et al. Exposed surfaces on shape-controlled ceria nanoparticles revealed through AC-TEM and water–gas shift reactivity. Chemsuschem 2013, 6, 1898–1906. [Google Scholar] [CrossRef] [PubMed]
  76. Dong, C.; Zho, Y.; Ta, N.; Shen, W. Formation mechanism and size control of ceria nanocubes. CrystEngComm 2020, 22, 3033–3041. [Google Scholar] [CrossRef]
  77. Li, C.; Luan, Y.; Zhao, B.; Kumbhar, A.; Zhang, F.; Fang, J. Shape Engineering of CeO2-Based Catalysts; Cambridge University Press: Cambridge, UK, 2020. [Google Scholar]
  78. Lykaki, M.; Sónia, E.P.; Carabineiro, C.; Andriopoutou, E.C.; Kallithrakas-Kontos, N.; Boghosian, S.; Konsolakis, M. Ceria nanoparticles shape effects on the structural defects and surface chemistry: Implications in CO oxidation by Cu/CeO2 catalysts. Appl. Catal. B Environ. 2018, 230, 18–28. [Google Scholar] [CrossRef]
  79. Russel-Webster, B.; Abboud, K.A.; Christou, G.C. Molecular particles of cerium dioxide structure-directing effect of halide ions. Chem. Commun. 2020, 56, 5382–5385. [Google Scholar] [CrossRef]
  80. Zhang, J.; Ohara, S.; Umetsu, M.; Naka, T.; Hakateyama, Y.; Adshiri, T. Colloidal Ceria Nanocrystals: A Tailor-Made Crystal Morphology in Supercritical Water. Adv. Mater. 2007, 19, 203–206. [Google Scholar] [CrossRef]
  81. Castanet, U.; Feral-Martin, C.; Demourgues, A.; Neale, R.L.; Sayle, D.C.; Caddeo, F.; Flitcroft, J.M.; Caygill, R.; Pointon, B.J.; Molinari, M.; et al. Controlling the {111}/{110} Surface Ratio of Cuboidal Ceria Nanoparticles. ACS Appl. Mater. Interfaces 2019, 11, 11384–11390. [Google Scholar] [CrossRef]
  82. Yan, H.; Liu, Z.; Yang, S.; Yu, X.; Liu, T.; Guo, Q.; Li, J.; Wang, R.; Peng, Q. Stable and Catalytically Active Shape-Engineered Cerium Oxide Nanorods by Controlled Doping of Aluminium Cations. ACS Appl. Mater. Interfaces 2020, 12, 37774–37783. [Google Scholar] [CrossRef]
  83. Bezkrovnyi, O.S.; Małecka, B.; Lisiecki, R.; Ostroushko, V.; Thomas, G.S.; Gorantla, A.; Kępinski, L. The effect of Eu doping on the growth, structure and red-ox activity of ceria nanocubes. CrystEngComm 2018, 20, 1698–1704. [Google Scholar] [CrossRef]
  84. Symington, A.R.; Molinari, M.; Moxon, S.; Flitcroft, J.M.; Sayle, D.C.; Parker, S.C. Strongly Bound Surface Water Affects the Shape Evolution of Cerium Oxide Nanoparticles. J. Phys. Chem. C 2020, 124, 3577–3588. [Google Scholar] [CrossRef]
  85. Breysse, M.; Guenin, M.; Claudel, B.; Latreille, H.; Veron, J. Catalysis of carbon monoxide oxidation by cerium dioxide: I. Correlations between catalytic activity and electrical conductivity. J. Catal. 1972, 27, 275–280. [Google Scholar] [CrossRef]
  86. Ta, N.; Liu, J.J.; Chenna, S.; Crozier, P.A.; Li, Y.; Chen, A.; Shen, W. Stabilized Gold Nanoparticles on Ceria Nanorods by Strong Interfacial Anchoring. J. Am. Chem. Soc. 2012, 134, 20585–20588. [Google Scholar] [CrossRef]
  87. Bezkrovnyi, O.S.; Kraszkiewicz, P.; Ptak, M.; Kępinski, L. Thermally induced reconstruction of ceria nanocubes into zigzag {111}-nanofacetted structures and its influence on catalytic activity in CO oxidation. Catal. Commun. 2018, 117, 94–98. [Google Scholar] [CrossRef]
  88. Yang, C.; Yu, X.; Heissler, S.; Nefedov, A.; Colussi, S.; Llorca, J.; Trovarelli, A.; Wang, Y.; Woll, C. Surface Faceting and Reconstruction of Ceria Nanoparticles. Angew. Chem. Int. Ed. 2017, 56, 375–379. [Google Scholar] [CrossRef]
  89. Soler, L.; Casanovas, A.; Ryan, J.; Angurell, I.; Escudero, C.; Pérez-Dieste, V.; Llorca, J. Dynamic Reorganization of Bimetallic Nanoparticles under Reaction Depending on the Support Nanoshape: The Case of RhPd over Ceria Nanocubes and Nanorods under Ethanol Steam Reforming. ACS Catal. 2019, 9, 3641–3647. [Google Scholar] [CrossRef]
  90. Li, J.; Liu, Z.; Cullen, D.A.; Wang, R. Ruthenium Diffusion on Different CeO2 Surfaces: Support Shape Effect. Microsc. Microanal. 2019, 25, 2198–2199. [Google Scholar] [CrossRef]
  91. Liu, S.; Wu, X.D.; Tamg, J.; Cui, P.Y.; Jiang, X.Q.; Chang, C.G.; Liu, W.; Gao, Y.X.; Li, M.; Weng, D. An exploration of soot oxidation over CeO2-ZrO2 nanocubes: Do more surface oxygen vacancies benefit the reaction? Catal. Today 2017, 281, 454–459. [Google Scholar] [CrossRef]
  92. Liu, Y.; McCue, A.J.; Yang, P.; He, Y.; Zheng, L.; Cao, X.; Man, Y.; Feng, J.; Anderson, J.A.; Li, D. Support-morphology dependant alloying behaviour and interfacial effects of bimetallic Ni-Cu/CeO2 catalysts. Chem. Sci. 2019, 10, 3556–3566. [Google Scholar] [CrossRef] [Green Version]
  93. Wang, F.; Zhang, L.; Zhu, J.; Han, B.; Zhao, L.; Yu, H.; Deng, Z.; Shi, W. Study on different CeO2 structure stability during ethanol steam reforming reaction over Ir/CeO2 nanocatalysts. Appl. Catal. A Gen. 2018, 564, 226–233. [Google Scholar] [CrossRef]
  94. Bezkrovnyj, O.S.; Kraszkiewicz, P.; Mista, P.; Kępinski, L. The Sintering of Au Nanoparticles on Flat {100}, {111} and Zigzagged {111}-Nanofacetted Structures of Ceria and Its Influence on Catalytic Activity in CO Oxidation and CO PROX. Catal. Lett. 2021, 151, 1080–1090. [Google Scholar] [CrossRef]
  95. Sartoretti, E.; Novara, C.; Fontana, M.; Giorgis, F.; Plumetti, M.; Bensaid, S.; Russo, N.; Fino, D. New insights on the defect sites evolution during CO oxidation over doped ceria nanocatalysts probed by in situ Raman spectroscopy. Appl. Catal. A Gen. 2020, 596, 117517. [Google Scholar] [CrossRef]
  96. Vanderspurt, T.H.; Wijzen, F.; Tang, X.; Leffler, M.P. Ceria-Based Mixed-Metal Oxide Structure, Including Method of Making and Use. U.S. Patent 0186805 A1, 28 March 2003. [Google Scholar]
  97. Susono, Y.T.; Toyota, M.M.; Susono, S.T.; Chiryu, N.T. High Hest-Resistant Catalyst with a Porous Ceria Support. U.S. Patent 5972830, 11 December 1995. [Google Scholar]
  98. Luccini, E.; Meriani, S.; Sbaizero, O. Preparation of zirconia-ceria powders by coprecipitation of a mixed zirconium carbonate in water with urea. Int. J. Mater. Prod. Technol. 1989, 4, 167–175. [Google Scholar]
  99. Si, R.; Zhang, Y.-W.; Li, S.-J.; Lin, B.-X.; Yan, C.-H. Urea-Based Hydrothermally Derived Homogeneous Nanostructured Ce1-xZrxO2 (x = 0−0.8) Solid Solutions: A Strong Correlation between Oxygen Storage Capacity and Lattice Strain. J. Phys. Chem. B. 2004, 108, 12481–12488. [Google Scholar] [CrossRef]
  100. Wolski, L.; Nowaczyk, G.; Jurga, S.; Ziolek, M. Influence of Co-Precipitation Agent on the Structure, Texture and Catalytic Activity of Au-CeO2 Catalysts in Low-Temperature Oxidation of Benzyl Alcohol. Catalysts 2021, 11, 641. [Google Scholar] [CrossRef]
  101. Mileva, A.; Issa, G.; Henych, J.; Štengl, V.; Kovacheva, D.; Tsoncheva, T. CeO2 and TiO2 obtained by urea assisted homogeneous hydrolyses method as catalysts for environmental protection: Effect of Ti/Ce ratio. Bulg. Chem. Commun. 2017, 49, 77–83. [Google Scholar]
  102. Pechini, M.P. Method of Preparing Lead and Alkaline Earth Titanates and Niobates and Coating Method Using the Same to form a Capacitor. U.S. Patent 3330697A, 11 July 1967. [Google Scholar]
  103. Dahl, J.A.; Maddux, L.; Hutchison, J.E. Toward Greener Nanosynthesis. Chem. Rev. 2007, 107, 2228–2269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Li, P.; Kumar, A.; Ma, J.; Kuang, Y.; Luo, L.; Sun, Y. Density gradient ultracentrifugation for colloidal nanostructures separation and investigation. Sci. Bull. 2018, 63, 645–662. [Google Scholar] [CrossRef] [PubMed]
  105. Danks, A.E.; Hall, S.R.; Schnepp, Z. The Evolution of ‘Sol-Gel’ Chemistry as a Technique for Materials Synthesis. Mater. Horiz. 2016, 3, 91–112. [Google Scholar] [CrossRef] [Green Version]
  106. Terrible, D.; Trovarelli, A.; Llorca, J.; Leitenburg, C.; Dolcetti, G. The preparation of high surface area CeO2-ZrO2 mixed oxides by a surfactant-assisted approach. Catal. Today 1998, 43, 79–88. [Google Scholar] [CrossRef]
  107. Chavhan, M.P.; Lu, C.-H.; Som, S. Urea and surfactant assisted hydrothermal growth of ceria nanoparticles. Coll. Surf. A Physicochem. Eng. Asp. 2020, 601, 124944. [Google Scholar] [CrossRef]
  108. Mori, K.; Jida, H.; Kuwahara, Y.; Yamashita, H. CoOx-decorated CeO2 heterostructures effects of morphology on their catalytic properties in diesel soot combustion. Nanoscale 2020, 12, 1779–1789. [Google Scholar] [CrossRef]
  109. Zhao, W.; Li, Z.; Wang, Y.; Fan, R.; Zhang, C.; Wang, Y.; Guo, X.; Wang, R.; Zhang, S. Ce and Zr Modified WO3-TiO2 Catalysts for Selective Catalytic Reduction of NOx by NH3. Catalysts 2018, 8, 375. [Google Scholar] [CrossRef] [Green Version]
  110. Reddy, B.M.; Lakshmanan, P.; Khan, A.; Loridant, S.; López-Cartes, C.; Rojas, T.C.; Fernández, A. Surface Stabilized Nanosized CexZr1-xO2 Solid Solutions over SiO2: Characterization by XRD, Raman, and HREM Techniques. J. Phys. Chem. B 2005, 109, 13545–13552. [Google Scholar] [CrossRef]
  111. Cargnello, M.; Doan-Nguyen, V.V.T.; Gordon, T.R.; Diaz, R.E.; Stach, E.A.; Gorte, R.J.; Fornasiero, P.; Murray, C.B. Control of metal nanocrystal size reveals metal-support interface role for ceria catalysts. Science 2013, 341, 771–773. [Google Scholar] [CrossRef] [Green Version]
  112. Ai, C.; Zhang, Y.; Wang, P.; Wang, W. Catalytic Combustion of Diesel Soot on Ce/Zr Series Catalysts Prepared by Sol-Gel Method. Catalysts 2019, 9, 646. [Google Scholar] [CrossRef] [Green Version]
  113. Neyertz, C.A.; Banús, E.D.; Miró, E.E.; Querini, C.A. Potassium-promoted Ce0.65Zr0.35O2 monolithic catalysts for diesel soot combustion. Chem. Eng. J. 2014, 248, 394–405. [Google Scholar] [CrossRef]
  114. Bolivar-Diaz, C.I.; Conesa, J.C.; Cortes-Corberan, V.; Monte, M.; Martinez-Arias, A. Nanostructured Catalysts Based on Combinations of Cobalt and Cerium Oxides for CO Oxidation and Effect of the Presence of Water. J. Nanosci. Nanotechnol. 2017, 17, 3816–3823. [Google Scholar] [CrossRef]
  115. Wright, C.S.; Walton, R.I.; Thompsett, D.; Fisher, J.; Ashbrook, S.E. One-Step Hydrothermal Synthesis of Nanocrystalline Ceria-Zirconia Mixed Oxides: The Beneficial Effect of Sodium Inclusion on Redox Properties. Adv. Mater. 2007, 19, 4500–4504. [Google Scholar] [CrossRef]
  116. Papadopoulos, C.; Kappis, K.; Papavasiliou, J.; Vakros, J.; Kuśmierz, M.; Gac, W.; Georgiou, Y.; Deligiannakis, Y.; Avgouropoulos, D. Copper-promoted ceria catalysts for CO oxidation reaction. Catal. Today 2020, 355, 647–653. [Google Scholar] [CrossRef]
  117. Xiao, G.; Li, S.; Li, H.; Chen, L. Synthesis of Doped Ceria with Mesoporous Flowerlike Morphology and Its Catalytic Performance for CO Oxidation. Micropor. Mesopor. Mater. 2009, 120, 426–431. [Google Scholar] [CrossRef]
  118. Si, R.; Zhang, Y.-W.; Wang, L.-M.; Li, S.-J.; Lin, B.-X.; Chu, W.-S.; Wu, Z.-Y.; Yan, C.-H. Enhanced Thermal Stability and Oxygen Storage Capacity for CexZr1-xO2 (x = 0.4−0.6) Solid Solutions by Hydrothermally Homogenous Doping of Trivalent Rare Earths. J. Phys. Chem. C 2007, 111, 787–794. [Google Scholar] [CrossRef]
  119. Zhang, Y.; Cui, M.; Wang, H.; Wang, L.; Hou, Y. Effects of ammonia concentration in hydrothermal treatment on structure and redox properties of cerium zirconium solid solution. J. Rare Earths 2021, 39, 419–426. [Google Scholar] [CrossRef]
  120. Zhu, Y.; Seong, G.; Noguchi, T.; Yoko, A.; Tomai, T.; Takami, S.; Adschiri, T. Highly Cr-Substituted CeO2 Nanoparticles Synthesized Using a Non-equilibrium Supercritical Hydrothermal Process: High Oxygen Storage Capacity Materials Designed for a Low-Temperature Bitumen Upgrading Process. ACS Appl. Energy Mater. 2020, 3, 4305–4319. [Google Scholar] [CrossRef]
  121. Devaraju, M.K.; Liu, X.; Yusuke, K.; Yin, S.; Sato, T. Solvothermal Synthesis and Characterization of Ceria-Zirconia Mixed Oxides for Catalytic Applications. Nanotechnology 2009, 20, 405606. [Google Scholar] [CrossRef]
  122. Burri, D.R.; Choi, K.-M.; Lee, J.-H.; Han, D.-S.; Park, S.-E. Influence of SBA-15 Support on CeO2-ZrO2 Catalyst for the Dehydrogenation of Ethylbenzene to Styrene with CO2. Catal. Commun. 2007, 8, 43–48. [Google Scholar] [CrossRef]
  123. Xin, J.; Cui, H.; Cheng, Z.; Zhou, Z. Bimetallic Ni-Co/SBA-15 catalysts prepared by urea co-precipitation for dry reforming of methane. Appl. Catal. A Gen. 2018, 554, 95–104. [Google Scholar] [CrossRef]
  124. Bao, J.; Chen, H.; Yang, S.; Zhang, P. Mechanochemical redox-based synthesis of highly porous CoxMn1-xOy catalysts for total oxidation. Chinese J. Catal. 2020, 41, 1791–1811. [Google Scholar] [CrossRef]
  125. Arandiyan, H.; Dai, H.; Ji, K.; Sun, H.; Li, J. Pt Nanoparticles Embedded in Colloidal Crystal Template Derived 3D Ordered Macroporous Ce0.6Zr0.3Y0.1O2: Highly Efficient Catalysts for Methane Combustion. ACS Catal. 2015, 5, 1781–1793. [Google Scholar] [CrossRef]
  126. Zhang, D.S.; Du, X.J.; Shi, L.Y.; Gao, R.H. Shape-controlled synthesis and catalytic application of ceria nanomaterials. Dalton Trans. 2012, 41, 14455–14475. [Google Scholar] [CrossRef] [PubMed]
  127. Pan, C.; Zhang, D.; Shi, L. CTAB assisted hydrothermal synthesis, controlled conversion and CO oxidation properties of CeO2 nanoplates, nanotubes, and nanorods. J. Solid State Chem. 2008, 181, 1298–1306. [Google Scholar] [CrossRef]
  128. Fuentes, R.O.; Acuna, L.M.; Zimic, M.G.; Lamas, D.G.; Sacanell, J.G.; Leyva, A.G.; Baker, R.T. Formation and Structural Properties of Ce−Zr Mixed Oxide Nanotubes. Chem. Mater. 2008, 20, 7356–7363. [Google Scholar] [CrossRef]
  129. Rombi, E.; Cutrufello, M.G.; Atzori, L.; Monaci, R.; Ardu, A.; Gazzoli, D.; Deiana, P.; Ferino, I. CO Methanation on Ni-Ce Mixed Oxides Prepared by Hard Template Method. Appl. Catal. A Gen. 2016, 515, 144–153. [Google Scholar] [CrossRef]
  130. Li, H.; Zhang, L.; Dai, H.; He, H. Facile Synthesis and Unique Physicochemical Properties of Three-Dimensionally Ordered Macroporous Magnesium Oxide, γ-Alumina, and Ceria-Zirconia Solid Solutions with Crystalline Mesoporous Walls. Inorg. Chem. 2009, 48, 4421–4434. [Google Scholar] [CrossRef]
  131. Rood, S.C.; Ahmet, H.B.; Gomez-Ramon, A.; Torrente-Murciano, L.; Reina, T.R.; Eslava, S. Enhanced ceria nanoflakes using graphene oxide as a sacrificial template for CO oxidation and dry reforming of methane. Appl. Catal. B Environ. 2019, 242, 358–368. [Google Scholar] [CrossRef]
  132. Han, W.Q.; Wu, L.J.; Zhu, Y.M. Formation and Oxidation State of CeO2-x Nanotubes. J. Am. Chem. Soc. 2005, 127, 12814–12815. [Google Scholar] [CrossRef]
  133. Tang, C.C.; Bando, Y.; Liu, B.D.; Goldberg, D. Cerium Oxide Nanotubes Prepared from Cerium Hydroxide Nanotubes. Adv. Mater. 2005, 17, 3005–3009. [Google Scholar] [CrossRef]
  134. Zhou, K.B.; Yang, Z.Q.; Yang, S. Highly Reducible CeO2 Nanotubes. Chem. Mater. 2007, 19, 1215–1217. [Google Scholar] [CrossRef]
  135. Chen, G.Z.; Xu, C.X.; Song, X.Y.; Xu, S.L.; Ding, Y.; Sun, S.X. Template-free Synthesis of Single-Crystalline-like CeO2 Hollow Nanocubes. Cryst. Growth Des. 2008, 8, 4449–4453. [Google Scholar]
  136. Reddy, L.H.; Devaiah, D.; Reddy, B.M. Microwave Assisted Synthesis: A Versatile Tool for Process Intensification. In Industrial Catalysis and Separations; Raghavan, K.V., Reddy, B.M., Eds.; Apple Academic Press, Inc.: Palm Bay, FL, USA, 2014; Chapter 10; pp. 375–405. [Google Scholar]
  137. Reddy, B.M.; Bharali, P.; Seo, Y.-H.; Prasetyanto, E.A.; Park, S.-E. Surfactant-Controlled and Microwave-Assisted Synthesis of Highly Active CexZr1-xO2 Nano-Oxides for CO Oxidation. Catal. Lett. 2008, 126, 125–133. [Google Scholar]
  138. Bang, J.H.; Suslick, K.S. Applications of Ultrasound to the Synthesis of Nanostructured Materials. Adv. Mater. 2010, 22, 1039–1059. [Google Scholar] [CrossRef] [PubMed]
  139. Alammar, T.; Noei, H.; Wang, Y.; Grünert, W.; Mudring, A.-V. Ionic Liquid-Assisted Sonochemical Preparation of CeO2 Nanoparticles for CO Oxidation. ACS Sustain. Chem. Eng. 2015, 3, 42–54. [Google Scholar]
  140. Tinoco, M.; Fernandez-Garcia, S.; Lopez-Haro, M.; Hungria, A.B.; Chen, X.; Blanco, G.; Perez-Omil, J.A.; Collins, S.E.; Okuno, H.; Calvino, J. Critical Influence of Nanofaceting on the Preparation and Performance of Supported Gold Catalysts. J. ACS Catal. 2015, 5, 3504–3513. [Google Scholar]
  141. Chen, W.; Li, F.; Yu, J.; Liu, L.; Gao, H. Rapid Synthesis of Mesoporous Ceria-Zirconia Solid Solutions via a Novel Salt-Assisted Combustion Process. Mater. Res. Bull. 2006, 41, 2318–2324. [Google Scholar]
  142. Baneshi, J.; Haghighi, M.; Ajamein, H.; Abdollahifar, M. Homogeneous precipitation and urea-nitrate combustion preparation of nanostructured CuO/CeO2/ZrO2/Al2O3 oxides used in hydrogen production from methanol for fuel cells. Part. Sci. Technol. 2020, 38, 464–474. [Google Scholar] [CrossRef]
  143. Kim, M.; Laine, R.M. One-Step Synthesis of Core-Shell (Ce0.7Zr0.3O2)x (Al2O0.3)1-x [(Ce0.7Zr0.3O2)@Al2O3] Nanopowders via Liquid-Feed Flame Spray Pyrolysis (LF-FSP). J. Am. Chem. Soc. 2009, 131, 9220–9229. [Google Scholar] [CrossRef] [PubMed]
  144. Strobel, R.; Pratsinis, S.E. Effect of Solvent Composition on Oxide Morphology during Flame Spray Pyrolysis of Metal Nitrates. Phys. Chem. Chem. Phys. 2011, 13, 9246–9252. [Google Scholar] [CrossRef]
  145. Sun, W.; Li, X.; Sun, C.; Huang, Z.; Xu, H.; Shen, W. Insights into the Pyrolysis Processes of Ce-MOFs for Preparing Highly Active Catalysts of Toluene Combustion. Catalysts 2019, 9, 682. [Google Scholar] [CrossRef] [Green Version]
  146. McCormick, P.G.; Tsuzuki, T.; Robinson, J.S.; Ding, J. Nanopowders synthesized by mechanochemical processing. Adv. Mater. 2002, 13, 1008–1010. [Google Scholar] [CrossRef]
  147. Sepelik, V.; Duvel, A.; Wilkening, M.; Becker, K.-D.; Heitjans, P. Mechanochemical reactions and synthesis of oxides. Chem. Soc. Rev. 2013, 42, 7507–7520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Boldyreva, E.V. Mechanochemistry of inorganic and organic systems: What is similar? what is different? Chem. Soc. Rev. 2013, 42, 7719–7738. [Google Scholar] [PubMed]
  149. Tsuzuki, T. Mechanochemical synthesis of metal oxide particles. Commun. Chem. 2021, 4, 143. [Google Scholar] [CrossRef]
  150. Gomollon-Bell, F. Ten chemical innovations that will change our world: IUPAC identifies emerging technologies in chemistry in potential to make our planet more sustainable. Chem. Int. 2019, 41, 12–17. [Google Scholar]
  151. Michalchuk, A.L.; Boldyreva, E.V.; Belenguer, A.M.; Emmerling, F.; Boldyrev, V.V. Tribochemistry, mechanical alloing, mechanochemistry: What is in a name? Front. Chem. 2021, 9, 685789. [Google Scholar] [CrossRef] [PubMed]
  152. Trovarelli, A.; Zamar, F.; Llorca, J.; De Leitenburg, C.; Dolcetti, G.; Kiss, J.T. Nanophase Fluorite-Structured CeO2-ZrO2 Catalysts Prepared by High-Energy Mechanical Milling. J. Catal. 1997, 169, 490–502. [Google Scholar] [CrossRef]
  153. Cutrufello, M.G.; Ferino, I.; Solinas, V.; Primavera, A.; Trovarelli, A.; Auroux, A.; Picciau, C. Acid-Base Properties and Catalytic Activity of Nanophase Ceria-Zirconia Catalysts for 4-Methylpentan-2-Ol Dehydration. Phys. Chem. Chem. Phys. 1999, 1, 3369–3375. [Google Scholar] [CrossRef]
  154. Dodd, A.; Mccormick, P.; Tsuzuki, T. Nanocrystalline zirconia powders synthesized by mechanochemical processing. Mater. Sci. Eng. A 2001, A301, 54–58. [Google Scholar] [CrossRef]
  155. Zheng, Y.; Hu, Z.; Huang, H.; Ji, W.; Sun, M.; Chen, C. Synthesis and Characterization of Nanometer Ce0.75Zr0.25O2 Powders by Solid-State Chemical Reaction Method. J. Nanomater. 2011, 2011, 657516. [Google Scholar] [CrossRef] [Green Version]
  156. Tsusuki, T.; McCormick, P.G. Synthesis of ultrafine ceria powders by mechanochemical processing. J. Am. Ceram. Soc. 2001, 84, 1453–1458. [Google Scholar]
  157. Li, Y.X.; Chen, W.F.; Zhou, X.Z.; Gu, Z.Y.; Chen, C.M. Synthesis of CeO2 nanoparticles by mechanochemical processing and the inhibiting action of NaCl on particle agglomeration. Mater. Lett. 2005, 59, 48–52. [Google Scholar]
  158. Li, F.; Yu, X.; Pan, H.; Wang, M.; Xin, X. Synthesis of MO2 (M: Si, Ce, Sn) nanoparticles by solid-state reactions at ambient temperature. Solid State Ion. 2000, 2, 267–772. [Google Scholar]
  159. Shu, Y.; Chen, H.; Chen, N.; Duan, X.; Zhang, P.; Yang, S.; Bao, Z.; Wu, Z.; Dai, S. A principle for highly active metal oxide catalysts via NaCl-based solid solutions. Chem 2020, 6, 1723–1741. [Google Scholar]
  160. Cheng, H.; Tan, J.; Ren, Y.; Zhao, M.; Liu, J.; Wang, H.; Liu, J.; Zhao, Z. Mechanochemical Synthesis of Highly Porous CeMnOx Catalyst for the Removal of NOx. Ind. Eng. Chem. Res. 2019, 58, 16472–16478. [Google Scholar]
  161. Lucentini, I.; Serrano, I.; Soler, L.; Divins, N.J.; Llorca, J. Ammonia decomposition over 3D-printed CeO2 structures loaded with Ni. Appl. Catal. A Gen. 2020, 591, 117382. [Google Scholar] [CrossRef]
Table 1. Examples of ceria-based catalysts applied for various chemical processes.
Table 1. Examples of ceria-based catalysts applied for various chemical processes.
ReactionCatalystRef.
CO oxidation • RuO2/CeO2 with ultra-large mesopores[1]
• Au nanoparticles supported over nanoshaped CeO2[33]
• CuO/CeO2—different shaped nanocatalysts[68]
• CuO-CeO2-ZrO2—flower-like morphology[70]
• Au nanoparticles on {100}, {111} and {111} nanofacetted structures of CeO2[94]
• CeO2 catalysts[85]
• Thermally reconstructed CeO2 nanocubes[87]
• CeO2 nanocatalyst[95]
• Nanostructured CoxCe1−xOy [114]
• Cu-promoted CeO2 catalysts[116]
• Doped CeO2—mesoporous flower-like morphology[117]
• CeO2 nanoplates, nanotubes and nanorods[127]
• CeO2 nanoflakes—graphene oxide as template[131]
• Nanostructured CexZr1−xO2[137]
• Sonochemically prepared CeO2 nanoparticles[139]
Diesel soot oxidation • CeO2-ZrO2 nanocubes[91]
• CoOx-decorated CeO2 heterostructures[108]
• CeO2/ZrO2 prepared by sol–gel method[112]
• K-promoted Ce0.65Zr0.35O2 monolithic catalysts[113]
Ethanol steam reforming • Rh-Pd/CeO2 nanocubes and nanorods[89]
• Ir/CeO2 nanocatalysts[93]
• Nanostructured CuO/CeO2/ZrO2/Al2O3[142]
Dry reforming of methane • MOF-derived Ni/CeO2 catalyst[21]
• CeO2 nanoflakes—graphene oxide as sacrificial template[131]
Low-temperature methanol combustion • CoxCe1−xOy nanocatalysts[3]
• Nanocomposite catalysts MnxCe1−xO2/γ-Al2O3[5]
• Nanometric Ce1−x CuxO(2−δ) materials[12,24,67]
• Nanowire hierarchical CeO2 micro-spheres[6]
• Catalysts prepared by Ce-MOFs pyrolysis[145]
Methane combustion • Macroporous Ce0.6Zr0.3Y0.1O2[125]
Combustion of polycyclic aromatic hydrocarbons • Nanostructured CeO2[16]
Fast pyrolysis of organic vapors • Hierarchically structured CeO2[7]
HCl oxidation with O2 to obtain Cl2 • Mesoporous CeO2[15]
NO reduction with CO • Oxidized and reduced NiOx/CeO2[50]
NH3-SCR • Highly porous CeMnOx[160]
NH3 decomposition • 3D-printed CeO2 structures loaded with Ni[161]
Carboxylative coupling using CO2 • Mesoporous CeO2-supported Ag[2]
Dehydrogenation of ethylbenzene to styrene with CO2 • CeO2-ZrO2/SBA-15[122]
4-methylpentan-2-ol dehydration • Nanophase CeO2-ZrO2 nanocubes[153]
Oxidation of benzyl alcohol with TBHP • Mesoporous CeO2[4]
Oxidation of benzyl alcohol with O2 • Au-CeO2 catalysts[100]
Low-temperature bitumen upgrading • Cr-substituted CeO2 nanoparticles[120]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dziembaj, R.; Molenda, M.; Chmielarz, L. Synthesis and Specific Properties of the Ceria and Ceria-Zirconia Nanocrystals and Their Aggregates Showing Outstanding Catalytic Activity in Redox Reactions—A Review. Catalysts 2023, 13, 1165. https://doi.org/10.3390/catal13081165

AMA Style

Dziembaj R, Molenda M, Chmielarz L. Synthesis and Specific Properties of the Ceria and Ceria-Zirconia Nanocrystals and Their Aggregates Showing Outstanding Catalytic Activity in Redox Reactions—A Review. Catalysts. 2023; 13(8):1165. https://doi.org/10.3390/catal13081165

Chicago/Turabian Style

Dziembaj, Roman, Marcin Molenda, and Lucjan Chmielarz. 2023. "Synthesis and Specific Properties of the Ceria and Ceria-Zirconia Nanocrystals and Their Aggregates Showing Outstanding Catalytic Activity in Redox Reactions—A Review" Catalysts 13, no. 8: 1165. https://doi.org/10.3390/catal13081165

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

Article Metrics

Back to TopTop