Next Article in Journal
An Improved Method to Encapsulate Laccase from Trametes versicolor with Enhanced Stability and Catalytic Activity
Next Article in Special Issue
Editorial: Special Issue “New Concepts in Oxidation Processes”
Previous Article in Journal
Photocatalytic Membrane Reactor (PMR) for Virus Removal in Drinking Water: Effect of Humic Acid
Previous Article in Special Issue
Molecular Orientations Change Reaction Kinetics and Mechanism: A Review on Catalytic Alcohol Oxidation in Gas Phase and Liquid Phase on Size-Controlled Pt Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ag/CeO2 Composites for Catalytic Abatement of CO, Soot and VOCs

1
Laboratory of catalytic research, Tomsk State University, 36, Lenin Ave., Tomsk 634050, Russia
2
Institute for the Study of Nanostructured Materials, Via Ugo La Malfa 153, 90146 Palermo, Italy
*
Authors to whom correspondence should be addressed.
Catalysts 2018, 8(7), 285; https://doi.org/10.3390/catal8070285
Submission received: 23 June 2018 / Revised: 10 July 2018 / Accepted: 11 July 2018 / Published: 16 July 2018
(This article belongs to the Special Issue New Concepts in Oxidation Processes)

Abstract

:
Nowadays catalytic technologies are widely used to purify indoor and outdoor air from harmful compounds. Recently, Ag–CeO2 composites have found various applications in catalysis due to distinctive physical-chemical properties and relatively low costs as compared to those based on other noble metals. Currently, metal–support interaction is considered the key factor that determines high catalytic performance of silver–ceria composites. Despite thorough investigations, several questions remain debating. Among such issues, there are (1) morphology and size effects of both Ag and CeO2 particles, including their defective structure, (2) chemical and charge state of silver, (3) charge transfer between silver and ceria, (4) role of oxygen vacancies, (5) reducibility of support and the catalyst on the basis thereof. In this review, we consider recent advances and trends on the role of silver–ceria interactions in catalytic performance of Ag/CeO2 composites in low-temperature CO oxidation, soot oxidation, and volatile organic compounds (VOCs) abatement. Promising photo- and electrocatalytic applications of Ag/CeO2 composites are also discussed.

1. Introduction

Air pollution is a major environmental problem. According to the World health organization, ambient air pollution contributes to 6.7 percent of all deaths worldwide [1], and the emissions of harmful compounds from industrial plants and motor vehicles in crowded urban areas are getting more attention. By reducing the level of air pollution, countries can reduce the morbidity rates of heart disease, lung cancer, chronic and acute respiratory diseases, etc. Many substances cause air pollution, including carbon monoxide (CO), particulate matter, ozone, nitrogen dioxide, soot, sulfur dioxide, organic dyes, etc., with CO being the most common among these pollutants. Volatile organic compounds (VOCs) comprising organic compounds with an initial boiling point inferior or equal to 250 °C (measured at a standard pressure of 101.3 kPa) also impact pollution of indoor and outdoor air [2]. In a recent review [3], the authors consider several main classes of VOCs, including halogenated VOCs, aldehydes, aromatic compounds, alcohols, ketones, polycyclic aromatic hydrocarbons, etc.
Therefore, air cleaning is a pivotal challenge, and new solutions are required. Catalytic total oxidation of organic pollutants into CO2 and water is the most effective way to address this challenge. Metal/ceria-based catalysts were found promising heterogeneous catalysts for CO, soot and VOCs oxidation, and the highly dispersed noble metals (Me = Au, Pt, Pd, Ru, etc.) were used as the active components of these catalysts. The ceria-supported catalysts containing Pd [4,5,6,7,8,9,10,11], Pt [12,13,14,15,16], Au [17,18,19,20,21,22,23], Ru [24,25], Rh [26,27,28,29], and Cu [30,31] were proposed. Metal oxide-based catalysts (Co3O4 [32], MnOx [33], etc.) also attracted wide interest. However, a significant part of the developed catalysts has limited use under real conditions due to high costs of noble metals (the loading of palladium, platinum, gold is of 2–10 wt. %), relatively low stability to “hard” conditions of oxidation processes causing the loss of active component and reduction of the catalyst activity and selectivity. Therefore, the development of high-performance affordable and stable catalysts for low-temperature total oxidation of harmful compounds is still challenging. Efficiency and costs of such catalysts are connected with proper selection of the type of active component, support, and preparation method [34].
Recently, supported silver catalysts have brought about wide interest due to their high activity in low-temperature oxidation processes. Different supports are studied (SiO2, CeO2, MnOx, TiO2, Al2O3, ZrO2, etc.) [35,36,37,38]. It is shown that an enhanced catalytic activity of Ag-based catalysts can be achieved by using reducible metal oxides as supports and by controlling the metal–support interaction to provide synergistic effect between active sites of the support and noble metal [39].
Among the supports mentioned, CeO2 brings about high interest, since it combines exceptional redox and acid-base properties with oxygen storage, which can be controlled by proper preparation methods and treatments. Moreover, these distinctive properties of ceria cause its wide applications as a support for catalysts. For this reason, highly active and relatively inexpensive Ag/CeO2 composite is considered promising heterogeneous catalyst for total oxidation of harmful organic compounds, including formaldehyde [40], CO [41,42,43], soot [44,45,46,47,48,49]. The Ag/CeO2 composites can also be used in photo- [50] and electrocatalysis [51,52], reduction of NOx [41], methane oxidation reaction [53], preferential oxidation of CO in excess of H2 (PROX CO) [54,55] as well as in biochemistry due to the bactericidal properties of both ceria and silver [56]. It is noteworthy also that these composites are applied in selective oxidation of organic compounds.
In this review, we provide a survey of the current state of catalytic total oxidation of CO, soot, and VOCs over Ag/CeO2 catalysts. The reactions under consideration are discussed from the perspective of (1) morphology and size effects of both Ag and CeO2 particles, including their defective structure, (2) chemical and charge state of silver, (3) charge transfer between silver and ceria, (4) role of oxygen vacancies, (5) reducibility of support and the catalyst on the basis thereof.

2. Topical Processes

2.1. CO Oxidation

CO oxidation is one of the most studied reactions in catalysis science. It is of great fundamental and practical importance, since CO is formed as a by-product in many industrially important oxidation reactions (e.g., methanol oxidation to formaldehyde [57], ethylene glycol oxidation to glyoxal [58], etc.). CO seriously affects the environment and human health [59].
Ceria-based catalysts are among the most promising materials for CO oxidation [59,60,61]. A comparison of ceria-supported noble metals shown that Pd/CeO2 and Au/CeO2 catalysts were more active in CO oxidation than Ag/CeO2 [62,63]. However, the relatively low activity of Ag-containing catalysts in this case may be connected with non-optimal conditions of preparation and pre-treatment of Ag/CeO2 catalyst. Ag/SiO2 catalysts are known to be able to catalyze low-temperature CO oxidation even at temperatures below 0 °C [64,65,66,67]. Such factors as the size of Ag nanoparticles, the pre-treatment conditions of both support and catalyst, metal–support interaction determine the catalytic activity of silver catalysts in CO oxidation. In Ref. [68] it was shown that addition of CeO2 to Ag/SiO2 improved the catalytic activity in CO oxidation due to the cooperation of oxidative species on Ag and ceria. Thus, the study of Ag/CeO2 catalysts deserves special attention to reveal the reasons for high catalytic activity and find the approaches to its regulation.
According to literature, the method of Ag/CeO2 synthesis determines the catalytic properties. Thus, in Ref. [41] the 10% Ag/CeO2 catalysts were prepared by impregnation and deposition–precipitation techniques. The catalysts prepared by impregnation demonstrated higher activity in CO and propylene oxidation. This finding was associated with formation of Ag2+ species in these catalysts, confirmed by Electron Paramagnetic Resonance (EPR). Such species improve the redox properties due to creation of three different redox couples: Ag2+/Ag+, Ag2+/Ag0, and Ag+/Ag0.
The effect of shape of ceria nanoparticles on the catalytic properties of ceria-based catalysts is also discussed in the review [69]. In Ref. [70] synthesis of ceria nanopolyhedra, nanorods, and nanocubes by a hydrothermal method is described (Figure 1). The oxygen storage capacity of CeO2 nanorods and nanocubes was attributed to both surface and bulk oxygen species. The lowest oxygen storage capacity for ceria nanopolyhedra was attributed to a predominance of (111) boundaries on the surface of particles with low reaction ability toward CO. Thus, the shape-selective synthetic strategy may be used for designing the catalysts with desired oxidative activity.
In Ref. [71] the catalytic activity of ceria rods, cubes and octahedra was studied in CO oxidation. The highest activity of ceria nanorods was attributed to a predominance of (110) and (100) surfaces, while the lowest activity of ceria octahedra was caused by a predominance of (111) surface. The activity of different surfaces also depends on the energy of oxygen vacancy formation, which is predicted to follow the reverse order of lattice oxygen reactivity: (110) < (100) < (111). Supporting of silver on the surface of ceria nanoparticles with different shapes by conventional incipient wetness impregnation followed by calcination at 500 °C led to creation of additional oxygen vacancies in ceria surface [43]. Ag nanoparticles were suggested to facilitate the formation of oxygen vacancies in ceria surface in a larger extent than in case of positively charged Agn+ clusters. Thus, Ag loading (1 and 3 wt. %) in Ag/CeO2 affects the amount of Ag0 and Agn+ clusters that yields different concentrations of surface oxygen vacancies and, hence, different activity in CO oxidation. Ag0 nanoparticles (NPs) promote the reducibility of surface lattice oxygen and catalytic activity of CeO2 in CO oxidation. The control of the shape of CeO2 may be used as a strategy to design the metal/CeO2 catalysts with reduced amounts of noble metals. An increase of the Ag content from 1 to up to 3 wt. % mitigates the difference in turnover frequency (TOF) CO for the composites based on nanocubes and nanorods that allows concluding on the need of coexistence of charged Agn+ species and reduced Ag0 NPs on the CeO2 surface to create an active catalyst.
The role of oxygen vacancies of Ag/CeO2 catalysts in CO oxidation is also discussed in Ref. [72]. Using Raman spectroscopy, it was shown that Ag promoted the formation of oxygen vacancies in ceria. This effect is pronounced, when CeO2 and Ag/CeO2 were reduced in CO/N2 atmosphere up to 300 °C (Figure 2a,b). Treatment in oxygen atmosphere leads to the decreased amount of oxygen vacancies (Figure 2c,d). Thus, the introduction of Ag into CeO2 promotes the activation of lattice oxygen of ceria and formation of oxygen vacancies that is the main reason for enhanced catalytic activity of Ag/CeO2 in CO oxidation.
The role of the shape of ceria nanoparticles in CO oxidation over Ag/CeO2 was also discussed in terms of the complex or hierarchical structure of ceria. The Ag-based catalysts supported on mesoporous CeO2 prepared by hard-template method and surfactant-template method was studied in CO oxidation in Ref. [42]. Mesoporous ceria was prepared by hard-template method using the SBA-15 material as a template, which was etched by NaOH. Hexadecyl trimethyl ammonium bromide (CTAB) was used as a classical soft template to synthesize ceria by surfactant-template method. Mesoporous ceria prepared by hard-template method was the preferable support for Ag catalysts, and total conversion of CO (200 mg catalyst, 1% CO, a gas flow of 30 mL/min) for this catalyst was achieved at 65 °C. High activity of this catalyst was attributed to oxygen vacancies in mesoporous CeO2 support, which stabilizes dispersed silver and facilitates the transfer of electrons from Ag to CeO2 via the Ag–CeO2 interface. However, one cannot exclude the participation of SiO2 used as a template to produce mesoporous CeO2 in formation of Ag-containing species highly reactive toward low-temperature CO oxidation.
In Ref. [73] Ag/CeO2 catalysts with the Ag loading from 5 to 20 wt. % were prepared by the HCl etching of CuO/CeO2/Ag2O mixed oxides followed by CuO removal. The formation of Ag nanoparticles inside the ultrafine nanoporous CeO2 with sizes of pore channels below 20 nm was observed after reduction by glucose in solution. The obtained composites also showed enhanced catalytic activity in CO oxidation in comparison with CeO2–Ag composite prepared by co-precipitation method, and the highest catalytic activity was observed for catalysts with 10 wt. % loading of Ag (T50% ≈ 130 °C, 1% CO and 10% O2, WHSV of 60,000 mL g−1 h−1).
The CeO2 mesoporous spheres with a diameter of ~100 nm and Ag catalysts on the basis thereof were synthesized in Ref. [74] (Figure 3). CeO2 mesoporous spheres were synthesized using glycol as a solvent with addition of C2H5COOH in an autoclave at 180 °C for 200 min. Ag NPs were prepared separately, and their dispersion in cyclohexane was stirred together with CeO2 mesoporous spheres. The catalysts were characterized by high surface area (216 m2/g) and regular morphology. Ag molar content was 10%. CO conversion achieved 96.5% at 70 °C (100 mL/min) and the enhanced catalytic performance in CO oxidation was attributed to the unique structure of ceria support.
The catalysts with core-shell and yolk–shell structures also attract attention [75,76]. The Ag@CeO2 catalysts with a core-shell structure were prepared by surfactant-free method with subsequent annealing redox reaction between silver and ceria precursor during co-deposition [77]. The particles with metallic Ag cores with a diameter of 50–100 nm CeO2 shell with a thickness of 30–50 nm were tested in CO oxidation (catalyst mass was 100 mg, 1% CO, a gas flow of 20 mL/min). The calcination of Ag@CeO2 at 500 °C in air flow led to the growth of catalytic activity (100% CO conversion at ~120 °C) in comparison with freshly deposited precipitate and catalyst after hydrothermal treatment and drying at 80 °C. This growth of activity was attributed to the strengthened interfacial interactions between Ag core and CeO2 shell during the calcination process (confirmed by TPR-H2) and to the fast desorption of CO2 from the surface of catalyst that was shown by Fourier Transform Infrared (FTIR) spectroscopy of adsorbed CO2. The charge transfer due to enhanced metal–support interaction from Ag to CeO2 was shown by XPS [39]. It is noteworthy that one-, two- and three-coordinated OH groups were shown to exist over CeO2 surface [78], and their effect cannot be neglected.
Thus, according to the literature, Ag/CeO2 composites are promising catalysts for CO oxidation. The method of catalyst preparation, shape of ceria nanoparticles, and morphology of ceria are the factors determining the catalytic properties of the composites. Special attention is given to oxygen vacancies, and their concentration depends on the shape of ceria particles, amount of silver and charge states of its clusters/nanoparticles as well as pre-treatment conditions. Certainly, the presence of silver on the surface of ceria promotes the formation of oxygen vacancies and facilitates the growth of catalytic activity in CO oxidation. The features of interfacial interaction also should be considered since the transfer of electronic density from silver NPs to ceria accompanies metal–support interaction in Ag/CeO2 catalysts. These phenomena may play a key role in oxidative catalysis [79,80], reduction of nitroarenes [81], photocatalysis [82]. Different synthetic strategies may be developed to synthesize Ag/CeO2 with high activity in CO oxidation and find real application in industrial or indoor air purification from CO and VOCs.

2.2. Soot Oxidation

Soot is an amorphous impure carbon formed during incomplete combustion of fuels and hydrocarbons in internal combustion engines, coal burning, power-plant boilers, etc. It is formed as a by-product impairing the normal operation of combustion engines by fouling of exhaust systems, generation of exhaust plumes, blocking the pipes, etc. [83]. Soot particles are harmful to the human respiratory system since they cannot be filtered by upper airways. Thus, the development of materials that prevent the harmful impact of soot on the environment and human health is an important research and technology challenge. The soot combustion of diesel exhaust particulate occurs at temperatures above 600 °C, while typical diesel engine exhaust temperatures are in the range of 200–500 °C [84,85]. Therefore, the decreasing of the temperature of soot combustion is the main requirement for catalysts in this reaction.
The contact between soot and catalyst plays a key role in solid–solid reactions, and the observed catalytic activity depends on the gas–solid–solid interaction [86]. The contact conditions between soot and catalyst determine the combustion performance. In the literature two types of catalyst–soot contact studies under laboratory conditions are proposed: tight contact (TC) and loose contact (LC) [85,86,87]. The LC mode comprises a mixing or shaking of the catalyst–soot mixture with a spatula providing conditions for contact between soot particle and catalyst similar to those over diesel filter. TC mode is achieved by milling (ball or mortar milling) of the mixture during several minutes. Compared to the LC mode, the TC mode is less representative of the real contact conditions but is required to better understand and discriminate the morphologies [86,88].
Many effective catalytic systems have been proposed for soot combustion and other oxidation reactions [83,89]. Due to their unique physical-chemical properties, especially high redox properties and the lability of lattice oxygen, ceria and ceria-based materials also show high catalytic activity in total oxidation reactions, and soot oxidation to carbon dioxide is not an exception. Ceria also possesses high oxygen storage capacity (OSC), which allows using the oxide not only as a support or modifying additive, but also as a catalyst for soot oxidation. A selection of CeO2-based catalysts for soot oxidation is presented in Table 1. In Ref. [90] the catalytic activity of pure ceria prepared by co-precipitation method was described. Precipitation of aqueous solution of HNO3 and Ce(NO3)3 was carried out using the 0.4 M NaOH solution and 0.4 M Na2CO3. Combustion temperature of pure oxide samples was achieved in the region of 445–560 °C. The acidification of cerium precursor at the stage of catalyst preparation improved the catalytic performance of the obtained materials. The sample prepared by precipitation method using HNO3/Ce(NO3)3 = 2 had the highest catalytic activity with Tm = 465 °C. It is noteworthy that the use of large amounts of alkali metals at the stage of synthesis may significantly influence on the morphology and defective structure of cerium oxide, which will impact on the observed catalytic activity [91].
Morphology is known to play an important role in solid–solid reactions, where the number of contact points is a crucial criterion of activity. In Ref. [92] three different morphologies of pure cerium oxide were studied in soot oxidation reaction. The materials comprised (1) ceria nanofibers that capture the soot particles in several contact points, while having low specific surface area (~4 m2/g), (2) solution combustion synthesis ceria having an uncontrolled morphology, but higher specific surface area (31 m2/g), and (3) three-dimensional self-assembled (SA) ceria stars having high specific surface area (105 m2/g) and highly available contact points. The latter showed the highest catalytic activity, and the temperature of soot oxidation reduced from 614 to up to 403 °C for TC and to up to 552 °C in case of LC (Figure 4).
Comparing to the morphologies in groups 1 and 2, the three-dimensional shape of SA stars may involve more of the soot cake layer that can be a reason for enhancement of the total number of contact points and higher catalytic activity (Figure 5). SA stars also keep their high intrinsic activity after aging.
A comparison of the catalytic performance of pure ceria with different morphology under LC conditions was carried in [60], and the results were compared to those reported in Refs. [46,93,94,95]. The activity was shown to decrease in the following order: nanorod > nanocube > fiber > flake, and the lowest temperature of complete combustion of 485 °C is observed for nanorod samples.
In Ref. [96] hydrothermal and solvothermal methods were used to prepare nanostructured ceria with different morphology (nanorod, nanoparticle, and flake). The nanorod sample showed the best catalytic activity (soot combustion temperatures for TC and LC modes were 368 and 500 °C, respectively) that was attributed to the maximal amount of adsorbed oxygen species on its surface. Moreover, the high specific surface area, determined by BET (Brunauer Hemmet Teller) method, was pointed out to have a positive effect in improving the activity under the LC mode. In Ref. [97] hydrothermal method was used to prepare conventional polycrystalline ceria and single-crystalline ceria nanorods and nanocubes. The obtained samples differ by the surface formed ((100) surfaces were typical for nanocubes, a mixture of (100), (110) and (111) surfaces for nanorods, while (111) surface was obtained for conventional polycrystalline ceria). More reactive exposed surfaces demonstrated higher catalytic activity and soot oxidation becomes a surface-dependent reaction. Soot, while located at the soot–ceria interface, can reduce ceria, and such surface becomes the source of active superoxide ions. The formation energy of a surface oxygen vacancy is considered important for activity enhancement.
According to Ref. [48], the redox properties of ceria are an important, but not the major factor for catalytic soot oxidation. A comparison of fluorite-type oxides CeO2, Pr6O11, CeO2–ZrO2, ZrO2 characterized by high oxygen capacity revealed that the reactivity rather than quantity of oxygen species involved in oxygen release/storage processes is a favorable factor for low-temperature soot oxidation. CeO2 was shown to be much more active in soot oxidation, than Pr6O11 and CeO2–ZrO2 that had higher OSC values than pure CeO2. Using the electron spin resonance (ESR) method it was demonstrated that the reason was connected with the ability of the CeO2 surface to generate superoxide ions (O2) that can rapidly react with neighboring carbon or recombine to yield O2.
Despite unique physical-chemical properties, it is often not feasible to use pure ceria, since a significant loss of specific surface may occur due to thermal sintering, deactivation of redox pair, reduction of OSC leading to deterioration of catalytic activity [98], etc. Even small sintering causes a large impact on the crystallite sizes and the presence of oxygen vacancies, which significantly reduces the catalytic activity. The presence of metal ions in the ceria lattice allows reducing the effects of sintering and loss of catalytic activity along with a significant increase of OSC [99,100].
Special attention should be paid to the effect of introduction of Ag into the CeO2 structure. Loading of Ag NPs on CeO2 improves the reactivity of CeO2 lattice oxygen toward soot oxidation. Kinetic studies showed [45] that lattice oxygen of ceria interacting with Ag NPs had similar reactivity to the one of lattice oxygen in Ag2O. Ag NPs enhance reducibility of ceria (which was also shown in [101] and was attributed to reverse spillover of oxygen atoms from the Ag–CeO2 boundary to the Ag NPs along with other possible interpretations), but not the reoxidation ability of reduced ceria surface by dioxygen. Silver can become an agent that allows rapid formation of Ox. In Ref. [102] using cyclic H2-TPR and Raman studies, it was shown that both dissociative adsorption of gaseous oxygen and migration of bulk oxygen of ceria can be facilitated by silver. This results in a rapid generation of atomic oxygen over silver, which under the TC mode can transfer onto soot particle and lead to catalytic oxidation reaction [103]. If not, its spillover onto the ceria surface occurs, and the oxygen transforms to Ox through 2O–O2–2O –2O2− over the oxygen vacancies [45,49,57,104,105]. On the other hand, silver is proposed to participate in the reverse transformation of O to O2 [105].
In Figure 6 an effect of silver loading on the catalytic activity of ceria in soot oxidation is represented. The temperature of soot combustion shifted from 668 °C in case of combustion of pure soot to 393 °C for СeO2 and to up to 345 °C for the case of Ag/СeO2 [48].
By comparing the onset temperature, Ti, of soot oxidation over various metal-loaded CeO2 with different loadings (Figure 7) it results that Ti can be lowered with an increase of Ag loading from 357 to 324 °C (20 wt %). On the contrary, loading other metals, such as Pd, Pt, and Rh, could not improve the activity. This result supports that superoxides activated over silver are the active species responsible for low-temperature soot oxidation.
The catalysts for soot combustion have two main drawbacks, i.e. poor soot/catalyst contact and restricted amount of active site. The promising composites should possess relatively low specific surface area and have no micropores and small mesopores, which will provide the presence of a maximal number of active sites on the external surface of the grain and will facilitate the effectiveness of catalyst performance. Various preparation technique can be used to create such active surfaces. While the impregnation method still can be used [48], the relatively simple and economically feasible co-precipitation technique is considered the major way to prepare Ag/CeO2 catalysts for soot oxidation [44,49,104,108]. As a result, an opportunity exists to design favorable structure to transfer/diffuse the activated oxygen species to reaction zones of the catalyst and promote better catalyst–soot contact.
Among the catalysts prepared by co-precipitation technique, special interest is devoted to those with the “rice-ball” core-shell structure [49,104] comprising metallic Ag particles in the core surrounded by CeO2 particles. These catalysts possess a unique agglomerated structure with a diameter of about 100 nm, where large Ag particles (30–40 nm) and a large interface between the Ag and CeO2 particles cause its excellent catalytic performance in soot oxidation due to this morphological compatibility (the oxidation proceeds below 300 °C).
A less common way to prepare Ag/CeO2 catalysts for soot oxidation is the electrospinning method [47]. CeO2 nanofibers with diameters of 241–253 nm were produced using this method (Figure 8).
The Ag/CeO2 and CeO2 fibrous catalysts calcined at 500 °C exhibited an improved catalytic performance in soot oxidation caused by their large pore sizes related to the macroporous characteristics of the porous structure in CeO2. Large surface areas of CeO2 and Ag metallic species can contribute to high soot oxidation activity (Figure 9).
In Ref. [106] it is pointed out that under oxygen-rich conditions the activity of Ag/CeO2 catalysts is caused by oxygen vacancies near Ag particles, while under oxygen-poor conditions it is controlled by bulk oxygen vacancies. The generation and transfer of active oxygen are affected by combinations of both types of oxygen vacancies.
The mechanism of soot oxidation over Ag/CeO2 composites is also debating. Soot oxidation is a solid–solid–gas reaction, and there are two points of views on the predominant reaction mechanisms of soot oxidation in the literature [48,109,110,111]. On one hand, soot oxidation is initiated by the surface-active oxygen (peroxide and superoxide (O and O2) species), which may be activated by the oxygen vacancies. From the other hand, surface active oxygen comes from the bulk by migration of lattice oxygen. Ref. [99,108] describes a mechanism of metal oxide catalyst participation in redox cycle, where metal is subjected to repeated oxidation and reduction according to the following reaction set:
Mred + Ogas → Moxd—Oads
Moxd—Oads + Cf → Mred + SOC
SOC → CO/CO2,
where Mred and Moxd–Oads represent the reduced and oxidized states of the catalyst, respectively; Ogas and Oads are gaseous O2 and surface adsorbed oxygen species, respectively; Cf denotes a carbon active site or free site on the carbon surface, and SOC represents a surface carbon-oxygen complex.
According to this mechanism, atomic Oads species is formed through dissociative adsorption of gas-phase oxygen on the metal oxide surface, and then attacks the reactive free carbon site Cf yielding an oxygen-containing active intermediate. CO/CO2 are formed through the reaction between the intermediate and either Oads or gas-phase O2. The authors [45,99,108] suggest that in this mechanism the surface adsorbed oxygen species play the key role in soot oxidation, in contrast to CO oxidation that occurs through the Mars-van Krevelen mechanism. However, some researchers consider that the second reaction mechanism is prevalent in soot oxidation over ceria-based catalysts under real conditions [48] (Figure 10).
In case of reverse CeO2–Ag catalyst [49], a synergistic effect of Ag and CeO2 particles causes adsorption of gas-phase O2 followed by formation of atomic oxygen species and the process is facilitated due to large Ag–CeO2 interface. The O species on the silver surface migrates to the surface of ceria particles through the interface and transforms into Onx− species (Figure 11). These atomic oxygen species exist in equilibrium during soot oxidation. Then the mobile active Onx− species migrates onto soot particle through the soot–ceria contact and completely oxidizes the soot into CO2.
Another important problem that occurs in particulate filters under real conditions is connected with the loss of contact between the catalyst and solid reactant (e.g., unreactive ash). In Ref. [104] the catalytic soot oxidation was shown to occur, when a physical barrier of ash deposit exists between the catalyst and the solid soot, and the reaction proceeds without a direct catalyst–soot contact or any external energy applied (Figure 12). A CeO2–Ag catalyst prepared by the co-precipitation and a Ag/CeO2 catalyst prepared by impregnation showed catalytic activity for remote oxidation of soot separated by the deposition of alumina or calcium sulfate, while CeO2 catalyst did not. The remote oxidation effect is extended to more than 50 μm for both the CeO2–Ag and Ag/CeO2 catalysts, with the highest effect over the former catalyst. Based on the results of the ESR experiments, a mechanism for the observed phenomenon was proposed, in which a superoxide ion (O2) generated on the catalyst surface first migrated to the ash surface and then to the soot particles and then subsequently oxidizes it.
In [112] several model Ag/CeO2 catalysts with uniform structures and diverse surface oxygen vacancy (VO–s) contents were prepared by solution combustion method, and the processes of their activation and deactivation were considered (Figure 13). The VO–s content, conditions of catalyst–soot contact and extra oxygen supplier were pointed out as the most important structural factors in the activity of soot oxidation catalysts. The dioxygen concentration in the reaction atmosphere was assumed to influence the VO–s content, while ceria reduction was mentioned to occur around the catalyst–soot contact points and did not take place in the presence of O2. Moderate amounts of VO–s were shown to boost the catalytic activity by generating more Oxn− species, while their excess yields O2− instead of O2 that hinders the process. The interfacial reduction of ceria and insufficient O2 delivery and regeneration were suggested to determine the catalyst performance. The deactivation can be postponed by noble metal addition, resulting in accelerated soot combustion over noble metal-containing catalysts.
Thus, the development of catalysts with a special state of the deposited phase, characterized by a strong metal–support interaction, makes it possible to stop the migration of the deposited particles of the active phase, preventing the process of thermal aging (sintering) of the catalyst, which is one of the main problems in the operation of catalytic systems for cleaning emissions of internal combustion engines, both gasoline and diesel. The synergistic effect of Ag/CeO2 catalysts is determined by high activity, stability and is achieved by decreasing the costs for use of expensive metals, e.g., platinum [113,114,115], with saving of efficiency in the processes of catalytic cleaning of emissions of internal combustion engines. In this way, Ag/CeO2 composites are considered promising catalysts for soot oxidation.

2.3. VOCs Abatement

VOCs are a large group of organic chemicals having high vapor pressure and low boiling point at atmospheric pressure (these include, but are not limited to aldehydes, alcohols, aromatic compounds, etc.). These properties cause evaporation or sublimation of these compounds from liquid or solid state and entering the indoor and outdoor air. VOCs are known to possess high toxicity, poison the atmosphere and have a negative impact on human health and the environment [34]. To date, numerous ways to solve the challenge of air pollution, such as combustion of wastes, biodegradation [116], adsorption [117], plasmochemical decomposition [118], photocatalytic oxidation [119], ozonation [120], etc., have been proposed. The main drawback of these methods is the high-energy consumption that may be accompanied by the formation of formaldehyde and CO as well as the complexity of regeneration of the active phase (bacteria, adsorbents, photocatalysts). Catalytic oxidation of VOCs to carbon dioxide and water are considered the most promising methods to control the emissions [121,122,123,124]. The use of catalysts allows carrying out VOCs oxidation at relatively low temperatures at complete conversion. As a rule, two main types of effective catalysts for total oxidation of VOCs are developed, including supported metals (e.g., Au, Pt, Pd, Ag) [125,126,127,128,129,130] and transition metal oxides (CeO2, MnO2, Co3O4) [130,131,132,133]. The combination of noble metal and transition metal oxide used as a support or modifier is promising to increase the effectiveness of catalytic composites [134,135,136].
Currently, Ag–CeO2 composites represent both scientific and practical interest as catalysts for VOCs abatement, in particular oxidation of formaldehyde, methanol, toluene, acetone, etc. A selection of literature data on Ag/CeO2 composites used in VOCs abatement is presented in Table 2. Several articles were published on formaldehyde oxidation over Ag/CeO2 catalysts [40,137,138,139]. One of the pioneer works in this field was carried out by S. Imamura et al. [140], who suggested using Ag/CeO2 as catalysts for formaldehyde oxidation. High activity of the Ag/CeO2 composite was suggested to be governed by high dispersion of active silver on CeO2 and easier removal of surface oxygen as compared to the one over individual Ag or CeO2 components. The authors pointed out that compared to other group VIII metals, silver is less expensive and more abundant and shows high activity and durability, when high temperatures are not required.
Thus, in Refs. [137,138] a comparison of Ag/CeO2 catalysts with Ag-containing catalysts supported on various supports and the those with different active components supported on CeO2 was considered. Catalytic activity toward formaldehyde oxidation was shown to strongly depend on the Ag particle size and dispersion and the amount of active oxygen species [137]. The 100% formaldehyde conversion was achieved above 125 °C. In Ref. [138] the defective sites of mesostructured CeO2 support prepared by pyrolysis of oxalate precursor were suggested to increase oxygen vacancies able to absorb and activate dioxygen, and highly dispersed silver particles promote this process. This allowed achieving the complete formaldehyde conversion at 100 °C and was accompanied by a strong synergistic interaction between active component and CeO2 support causing enhancement of redox capability of the catalyst.
L. Ma et al. [40] also pointed out the synergistic interaction between Ag and CeO2 that caused an activity enhancement of Ag/CeO2 nanosphere catalysts with average sizes around 80–100 nm composed of small particles with a crystallite size of 2–5 nm as compared to normal Ag/CeO2 particle catalysts prepared by conventional impregnation method. The complete formaldehyde conversion was achieved above 110 °C, which was also explained by the fact that surface chemisorbed oxygen can be easily formed on the Ag/CeO2 nanosphere catalysts. Silver facilitated oxygen activation, which was considered an important aspect of formaldehyde oxidation.
Similar idea was reported in Ref. [139], where the comparison of catalytic properties of Ag/CeO2 catalyst with different morphologies (nanorods, nanoparticles, and nanocubes) of CeO2 prepared by hydrothermal and impregnation method was carried out. The authors pointed out shape dependence of the chemical state of ceria-supported Ag NPs, with the catalysts supported on CeO2 nanorods showing the highest activity caused by the highest surface oxygen vacancy concentration, high low-temperature reducibility as well as existence of lattice oxygen species and lattice defects formed with the participation of both silver and ceria. The electronic silver–ceria interaction yielded Ag0 in Ag/CeO2 composites, and the Ag0/(Ag0 +Ag+) ratio was found the highest for the catalysts supported on ceria nanorods. These results show that the catalytic activity of Ag/CeO2 composites toward formaldehyde abatement can be regulated by engineering the proper shapes of CeO2 supports.
One of the main parameters that allows comparing the catalytic activity of different materials is a TOF. Table 2 presents the TOF values calculated by the authors. Unfortunately, the differences in calculation methods and absence of required experimental information in original papers did not allow comparing the activity of Ag/CeO2 materials correctly.
Besides formaldehyde, Ag/CeO2 catalysts were also used to oxidize other VOCs, e.g. methanol, toluene, acetone, and naphthalene [41,54,141,142,143]. In these articles, a comparison of catalysts prepared by different methods was represented. The authors attempted to determine the influence of the preparation method and structure of the catalyst on its catalytic activity. Thus, in Ref. [54] the properties of catalysts prepared by deposition–precipitation and co-precipitation methods were compared in total oxidation of methanol, acetone, and toluene. The catalysts prepared by co-precipitation method were revealed to be more active in oxidation reactions. Small crystallites of silver and ceria enhanced the mobility and reactivity of oxygen species over ceria surface, which participated in the said reactions through the Mars-van Krevelen mechanism. The reactivity of the VOCs changed in a row: methanol > acetone > toluene.
In Refs. [41,142] the comparison of the catalytic activity of M/CeO2 composites (M = Au, Cu, Ag) prepared by conventional wet impregnation and deposition–precipitation methods was carried out in propylene oxidation. It was shown that the Ag-containing catalyst prepared by conventional wet impregnation method possessed higher catalytic activity. In Ref. [142] the presence of silver in high oxidation state was considered responsible for high catalytic activity of Ag/CeO2 composites. Using EPR technique it was shown that this is connected with the presence of Ag2+ ions (isotopes 107Ag2+ and 109Ag2+ were detected) along with Ag+ and Ag0 in the Ag/CeO2-Imp sample, while this was not observed in case of Ag/CeO2-DP (Figure 14, A) [41].
In the presence of Ag2+ ions, a mobility of some oxygen species increases, which sets conditions for the formation of three redox couples (Ag2+/Ag+, Ag2+/Ag0, and Ag+/Ag0). Nitrate precursor decomposition with the participation of O2− of ceria lattice was considered a source of Ag2+ ions, while the regeneration of oxygen vacancy may occur either from nitrate or from gaseous oxygen:
( Ag + + NO 3 ) / ( O 2 Ce 4 + O 2 )
( Ag + + NO 2 ) / ( O 2 Ce 4 + O 2 )
( Ag 2 + + O 2 ) / ( O 2 Ce 4 + O 2 ) + NO 2
In Figure 14B the catalytic conversion of propylene over CeO2, Ag/CeO2-Imp and Ag/CeO2-DP is shown. Adding Ag to CeO2 enhanced the catalytic activity, moreover, the performance of the Imp catalyst was better than that for the DP. In order to evaluate the stability of the catalyst over time, the authors also presented both static (isothermic conditions at 175 °C) and dynamic (7 consecutive cycles vs temperature in the range from 50 to up to 300 °C) aging tests for the activity of the 10% Ag/CeO2 (Imp) sample in propene oxidation. Moreover, EPR studies were carried out for the samples before and after catalysis. It was stated that after catalysis the Ag2+ ions retained on the ceria surface. This allows formulating the key role of Ag2+/Ag+ and Ag2+/Ag0 redox couples as active species in propene oxidation over 10% Ag/CeO2 by prepared impregnation method.
S. Benaissa et al. [141] prepared a mesoporous CeO2 using nanocasting pathway with SBA-15 as a structural template and cerium nitrate as a CeO2 precursor and compared the properties of catalysts on the basis thereof prepared by wetness impregnation (WI), deposition–precipitation with urea (DPU) and impregnation–reduction with citrate (IRC) methods, with the latter being the most active and stable (the catalytic activity and selectivity did not significantly change after 50 h). The authors connected this with higher surface lattice oxygen mobility over this catalyst and with strong silver–mesoporous ceria interaction.
The authors [143] carried out isothermal naphthalene oxidation comparing the activity of catalysts with different Ag content (0.5–5 wt. %), with the sample containing 1 wt. % Ag being the most active one. This was explained by the balance between two factors: oxygen availability and oxygen regeneration capacity. Introduction of Ag to CeO2 was shown to increase both factors. Regeneration capacity was related to the number of oxygen vacancies in bulk ceria, and Ag facilitated the process by reverse spillover effect. Cex+ ions were suggested to be the main active sites. Impregnated silver was claimed to serve as a “pump” and increase bulk oxygen vacancies, while reducing the surface ones, which resulted in oxygen availability and determined the oxygen regeneration. Spillover effect was proposed to reduce the regeneration ability of active oxygen, when Ag loading is high, which was connected with lower concentration of surface oxygen vacancies.
Of particular interest is the approach to locate the Ag/CeO2 composition on the inert support, which is usually represented by alumina or silica [144,145]. Thus, H. Yang et al. [144] used 3DOM CeO2–Al2O3 as a support for Ag catalysts for toluene oxidation. This support was prepared using the Pluronic F127 (EO106PO70EO106) and PMMA as soft and hard templates, respectively. The obtained support showed high-quality 3DOM architecture with a diameter of macropores of 180–200 nm, where ordered mesopores with a diameter of 4–6 nm were formed on the skeletons of macropores. Such structure allowed producing the particles of active component with sizes of 3–4 nm that were evenly distributed on the catalyst surface. The 50% and 90% toluene conversion (1000 ppm) over 0.81Ag/3DOM 26.9CeO2–Al2O3 sample was achieved at 308 and 338 °C, respectively.
In Ref. [145] silica gel prepared by sol–gel method and subjected to hydrothermal treatment was used as a primary support. Ceria and then silver were supported onto silica gel using consecutive impregnation method. The activity of the obtained catalysts was studied in formaldehyde oxidation reaction. The author pointed out that the activity of Ag/CeO2/SiO2 catalysts was significantly higher than the one of Ag/SiO2 sample, which was attributed to synergetic action between silver and ceria. The results obtained for the silver catalyst with small amounts of ceria were not significantly inferior to silver supported over bulk ceria (Figure 15). Thus, the silica-supported ceria-modified silver catalyst can be used for formaldehyde oxidation.
To conclude, Ag/CeO2 catalysts are promising materials for VOCs abatement. Even though their activity is inferior to the one of catalysts based on noble metals, their use still represents wide interest due to lower costs. Moreover, the opportunity to increase their activity due to the application of various preparation methods as well as changing of Ag/Ce ratio forms the ground for future research in this field. It is noteworthy that in the literature there is no consensus on the effect of preparation method of Ag/CeO2 composites on their catalytic activity in VOCs abatement.

3. Ag/CeO2 Composites: Insights from Theory

Due to low amounts of silver that are usually used in the preparation of highly effective Ag/CeO2 composites for total oxidation of VOCs, soot, and CO, not all experimental techniques can provide a representation of silver–ceria interface and the ways it works in the said catalytic transformations. Thus, Ag/CeO2 composites have attracted the attention of theoretical chemists. Two main directions are considered: (1) adequate representation and modeling of regular and defective ceria surfaces [132,146,147,148,149,150,151], (2) systematic studies of the adsorption behavior of Ag clusters on ceria surfaces [152,153,154,155,156,157,158,159,160]. In the latter case, the structure of Ag–ceria interface is widely discussed, while the adsorption behavior of adsorbates over such composites and their roles in tuning the interfacial properties are modeled in a lesser extent [152].
Researchers point out several difficulties in terms of theoretical modeling of CeO2-based composites. These difficulties are as follows: (1) density functional theory (DFT) does not predict correctly the localized nature of Ce 4f states, (2) change of Ce oxidation state causes incorrect lattice parameters, (3) the calculation results strongly depend on the used methods and functionals, and the obtained energy values oscillate.
These issues were partially addressed by application of hybrid functionals [132,161,162] or DFT+U approach [152,157,163]. The latter is connected with the inclusion of U term for highly correlated Ce 4f electrons in reduced ceria providing partial occupancy of the corresponding atomic level and increasing the accuracy of modeling of the on-site Coulomb interactions in CeO2-based materials. The values for U are usually selected semiempirically. The formalism by Dudarev et al. [164] is usually used. A combination of local density approximation (LDA) and generalized gradient approximation (GGA) in periodic calculations is shown to adequately describe geometry and energy parameters [165] under this approach. However, it is noteworthy that the results of DFT+U calculations depend on many parameters (e.g., lattice constants), which requires special attention to their interpretation.
In Ref. [160] using LDA+U and GGA+U DFT approaches with different U values and periodic slab surface models, charge transfer was shown to occur from Ag to ceria with a concomitant reduction of one Ce surface atom of the top layer, and the transferred electron was localized on Ce atoms. For Ag-based systems, the most favorable adsorption site comprised three surface oxygen atoms. In Ref. [159] the studies of surface structures and electrophilic states of Ag adsorbed on CeO2(111) revealed that charge redistribution can be caused by local structural distortion effects. The distribution of charge was not uniform over the top O layer because of Ag clusters on the underlying O ions, which increased the ionic charge of the remaining O ions and decreased the effective cationic charge over Ce atoms bonded with uncovered O atoms. This also influenced back on the structure of Ag cluster. Silver clusters were shown to induce changes in the oxidation state of several Ce atoms located in the top layer (Ce4+ to Ce3+), which are accompanied by a charge flow from metal cluster to surface caused by electronegativity difference between Ag and O atoms [154].
In Ref. [158] charge redistribution during Ag adsorption was confirmed by construction of spin density isosurfaces and site projected density of states. The distortions of selected Ce–O distances were imposed to study the energetics of Ce4+ to Ce3+ reduction. Oxidation of Ag0 to Ag+ was assumed, while the probable formation of partially oxidized AgxOy species was not considered. Two nearest neighbor Ce3+ sites relative to Ag showed the highest Ag adsorption energy at O bridge sites, while three nearest neighbor Ce3+ sites showed the highest Ag adsorption at Ce bridge sites.
DFT calculations were carried out for ceria-supported 4-atom transition metal (including Ag) clusters in Ref. [155] and showed that the strength of metal–metal and metal–oxygen interactions depended on the hybridization of d-states of metal with p-states of oxygen as well as the occupation of antibonding Ag d-states. The interactions changed the itinerant f-states of cerium to localized ones, which created a lateral tensile strain in the top layer of Ce on the surface. It was suggested also that the structure of Ag cluster determined the number of cerium atoms in the localized Ce3+ oxidation state.
Combined experimental (XPS, STM) and theoretical (DFT+U) approaches were used to study the nucleation and growth of Ag nanoparticles deposited on stoichiometric and reduced thin CeO2 films grown on Pt(111) [157]. A direct electron transfer from Ag clusters and nanoparticles to ceria was reported, and its extent, as well as spin, localization depended on the level of theory used. Ag atoms or nanoparticles supported on stoichiometric CeO2 acted as electron donors and are subjected to spontaneous direct oxidation at the expense of ceria followed by reduction of Ce ions of the support. The energy costs to move single O atom from ceria toward adsorbed Ag nanoparticle was high, and reverse spillover of oxygen cannot be considered a favorable mechanism of ceria reduction.
Silver–ceria interaction is often compared with the one in Au/CeO2 and Cu/CeO2 systems. Due to relatively lower ionization potential, Ag and Cu show higher adsorption energies. Moreover, silver nanoparticles act as a platform for oxygen diffusion leading to partially oxidized Ag nanoparticles located on the surface of the partially reduced ceria [157]. To quantitatively explore the interactions between silver and ceria, a method is proposed utilizing the conversion of total adsorption energy into the interaction energy per Ag–O bond and measurement of a deviation of Ag–O–Ce bond angle from the angle of the sp3 orbital hybridization of O atom [153]. It is noteworthy that coordination number of O atom, although generally considered, is not included into the correlation, while in Ref. [156] multiple adsorption configurations are shown to exist over single adsorption sites for Ag/CeO2(100), and electron charge transfer occurs between the neutral silver atom and neighboring Ce4+ cation.
In Ref. [152] the reactivity of Ag-modified CeO2(111) surface used in soot combustion was considered. The interactions of stoichiometric and reduced CeO2 (111) surfaces with dioxygen, carbon clusters, isolated Ag atoms and silver clusters were studied using DFT+U approach. Carbonaceous species yielded oxygenated carbon moieties of reduced ceria. Peroxo and superoxo species are shown to form, when O2 is adsorbed over Ag cluster. The role of Ag atoms is to act as a donor, which, when oxidized, donate the valence electron to ceria yielding reduced Ce3+ ions. The presence of small Ag clusters mediates the formation of oxygen vacancies (Figure 16).
The vacancies possess stronger affinity with respect to oxygen as compared to silver that leads to refilling of the cavities with dioxygen. Co-presence of Ag clusters and reduced ceria lightens electron transfer and activation of dioxygen molecule. Silver atoms perform as alkali metal promoters to facilitate O2 to O2 transition that leads to the formation of reduced Ce3+ ions. However, partial oxidation of silver can take place in this case.
Despite thorough investigations, still there are several debating issues in the theoretical description of Ag/CeO2 composites. Among them are the mechanism of oxygen replenishing in the support, different behavior of CeO2 surfaces, adsorption of silver atoms over long and short O–O bridge sites, quantitative description of Ag–CeO2 interactions, etc.

4. Emerging Applications

4.1. Photocatalysis

The wide application of CeO2-based catalysts in oxidative catalysis is mainly attributed to intrinsic redox properties [166]. Conversely, the interest in using ceria in photocatalysis is much lower. This is connected with fast recombination of photoinduced electron–hole pairs and limited visible light adsorption capacity [167]. CeO2 is an n-type semiconductor with a relatively wide bandgap (Eg = 3.15–3.2 eV) [167,168]. On the other hand, CeO2 has emerged as a promising material for photocatalysis owing to its chemical stability and photocorrosion resistance [169]. Redox Ce4+↔Ce3+ transition is accompanied by oxygen vacancy formation, which has high importance for both oxidative catalysis and electron–hole separation/recombination in photocatalyst [170]. Thus, in Ref. [171] a mesoporous nanorod-like ceria prepared by microwave-assisted hydrolysis of Ce(NO3)3*6H2O in the presence of urea was characterized by significant shifts of adsorption to the visible region (a band gap of 2.75 eV) that was associated with the presence of Ce3+. The growth of temperature was also shown to result in significant reduction of the recombination of photogenerated electron–hole pairs. The increased photocatalytic activity in gas-phase oxidation of benzene, hexane, and acetone was found for the prepared mesoporous nanorod-like ceria due to these two phenomena. Thus, the shape of ceria nanoparticles and the presence of Ce3+ in the structure provided a growth of photocatalytic activity, including the one under visible light.
Various strategies are being developed to improve the photocatalytic properties of ceria-based materials: morphology control [172,173], doping by europium or yttrium [174,175], fabrication of heterojunctions [176], etc. Thus, in Ref. [172] the degradation of the azo dye acid orange 7 (AO7) under ultraviolet irradiation over hierarchical rose-flower-like CeO2 nanostructures (Figure 17) is studied. The synthesis of CeO2 sheets active under the visible light is described in Ref. [173].
Moreover, the fabrication of CeO2-based heterostructures is a more promising way to reduce the band gap and provide improved electron–hole separation due to charge transfer through the interfacial boundaries. Silver salts may be used in photocatalysis due to their semiconductors properties. Thus, Ag3PO4 are characterized by relatively small band gap (2.36–2.43 eV) [177], absorb visible light (has yellow color) and possess a good photocatalytic stability. In Ref. [178] the photocatalytic activity of new composite Ag3PO4/CeO2 in degradation of methylene blue and phenol under visible light and UV light irradiation was studied. The photocatalytic activity of the Ag3PO4/CeO2 composite was shown to be associated with the fast transfer and efficient separation of electron–hole pairs at the interfaces of two semiconductors (CeO2 and Ag3PO4). The stability of photocatalyst was demonstrated during five catalytic cycles.
The photocatalytic remediation of water polluted by some chemically stable azo dyes using Ag2CO3/CeO2 microcomposite under visible light irradiation was studied in Ref. [179]. The enhanced photocatalytic activity for the photodegradation of enrofloxacin in aqueous solutions over Ag2O/CeO2 composites under visible light irradiation was demonstrated in Ref. [167]. The composite was synthesized by an in situ loading of Ag2CO3 on CeO2 followed by thermal decomposition. The p-n heterojuction between two semiconductors provided efficient separation of photoinduced charges through the contact of semiconductors that was shown by photoluminescence spectra (Figure 18a). The formation of Ag nanoparticles was associated with photoreduction of Ag2O. The surface plasmon resonance (SPR) on Ag NPs may lead to the formation of electrons and holes in such a way that the electrons could migrate from Ag NPs to the conduction band (CB) of Ag2O (Figure 18b). Thus, Ag NPs may play a specific role in photocatalytic degradation of organic pollutants.
The same effect of photoreduction of silver compounds with the formation of Ag NPs was observed for Ag/AgCl–CeO2 catalysts [180]. The energy of hot electrons, generated on Ag NPs due to SPR, is between 1.0 and 4.0 eV [181], and these electrons could migrate to the CB of AgCl in such a way that the electrons and holes generated on CeO2 and Ag NPs would be efficiently separated. Thus, in composite photocatalysts the role of Ag NPs in visible light adsorption and separation of charges is high.
The decoration of ceria by metals (Au, Pt, Pd, Ag) provides growth of photocatalitic activity due to increased electron–hole separation and extended time of light response of semiconductors [170]. The three main phenomena of charge transfer are involved through metal–semiconductor interface: Schottky barrier (transfer of electrons from semiconductor to metal) (Figure 19a), metal SPR with transfer of charge from metal to semiconductor (Figure 19b) and metal SPR—local electric field (accompanied by recombination of electrons from metal and holes of semiconductors) (Figure 19c). The SPR for Ag NPs is observed generally near the wave-length of 400 nm, while adsorption of Au NPs is observed at 550 nm [181], which makes gold more attractive for photocatalysis [182,183]. However, the position of the absorption band of nanoparticles depends on many factors, including the size and shape of particles, interaction with surroundings. Thus, significant shift of SPR of Ag NPs from 400 nm to 480–500 nm is observed for Ag/CeO2 catalysts [184] that may be attributed to strong electronic metal–support interaction between Ag and CeO2. This provides an enhanced photocatalytic activity of Ag/CeO2 composites in the degradation of methylene blue under the simulated sunlight [50] or visible light [185]. According to [50], Ag acts as an acceptor of photoelectrons, and then the electron rapidly reacts with O2 yielding O2 that reduces the probability of recombination of electron–hole pairs. The correlation between the rate of degradation and amount of Ag NPs (active sites) was found. High stability and high recyclability of the Ag/CeO2 heterostructure catalysts was shown.
In Ref. [186], a photocatalytic degradation of Congo Red under UV light and visible light over three-dimensionally ordered macroporous (3DOM) Ag/CeO2–ZrO2 material was studied. It was shown that the SPR effect of Ag particles provides the adsorption of visible light and promotes separation of electrons and holes, reducing their recombination and improving the photocatalytic activity. The superior photocatalytic activity of Ag/CeO2/ZnO nanostructure was shown in degradation of azo dyes (methylene orange and methylene blue) and phenol solution under visible light irradiation was demonstrated in Ref. [187]. It was found that formation of oxygen vacancies led to a narrow band gap (2.66 eV), which helps to produce sufficient electrons and holes under visible light in the ternary Ag/CeO2/ZnO nanostructure. The defect structure of composite inhibited the electron–hole recombination and provided synergistic effect of narrow band gap. The SPR of Ag NPs and defects (Ce3+ and oxygen vacancies) in CeO2 and ZnO resulted in superior photocatalytic activity. In Ref. [188], the correlation between Ce3+ loading, amount of oxygen vacancies and activity of Ag/CeO2 and Au/CeO2 catalysts in photodegradation of rhodamine blue dye in an aqueous medium under UV–vis irradiations were found. The conditions of synthesis (pH of precipitation) and Ag/Au loading provided different Ce3+ loading, distortion of CeO2 lattice and concentration of vacancies. All these parameters affected on light absorbance, separation of photogenerated charges and photocatalytic properties.
Thus, silver and its compounds supported on ceria have high importance in photocatalytic degradation of organic pollutants. Semiconductor properties of silver compounds and SPR of Ag NPs provide both absorbance of visible light, separation of electrons and holes and result in increased photocatalytic activity. Several common aspects were found between classical oxidation catalysis and photocatalysis over Ag/CeO2 composites. The interaction of silver with ceria (including electronic metal–support interaction) influences on the catalytic activity of Ag/CeO2 due to cooperation of active sites of Ag and ceria. The presence of Ag–CeO2 contact also leads to a growth of the amount of oxygen vacancies in the structure of CeO2 that also promotes an enhanced catalytic/photocatalytic activity. Generally, Ag/CeO2 composites are new for photocatalysis and poorly described. The study of Ag/CeO2 systems in photocatalysis has high importance for fundamental research and real application of catalysts in the purification of aqueous wastes from dyes and other organic pollutants.

4.2. Electrocatalysis

Silver was also shown to be a promising material for electrocatalytic applications [189,190,191]. Recently, ceria has attracted a growing interest as a component of materials for electrocatalytic applications [192,193]. The main reasons for this are its high oxygen storage and transfer abilities. Application of proper amounts of noble metal improves the conductive properties of CeO2-based materials, thus making them promising composites for electrocatalytic applications in fuel cells, metal-air batteries, and other alternative energy transfer devices [194].
A combination of silver and ceria in Ag/CeO2 composites was used in several publications [51,52,195,196]. In Ref. [196] the Ag/CeO2 composites comprising 30–50 nm silver nanoparticles uniformly anchored on the surface of nanosheet-constructing porous CeO2 microspheres were used as oxygen reduction reaction catalysts. CeO2 is known to show high oxygen storage capability and oxygen transfer ability, and silver was added to improve the conductivity of the latter. As a result, an enhanced activity was observed, and aluminum–air batteries based on Ag/CeO2 composites exhibited an output power density of 345 mW/cm2 and low degradation rate of 2.6% per 100 h, respectively.
In Ref. [51] a method was developed to prepare nanoporous Ag–CeO2 ribbons with a homogeneous pore/grain structure by dealloying melt-spun Al–Ag–Ce alloy in a 5 wt. % NaOH aqueous solution. The resulting structure comprised uniform CeO2 particles dispersed on the fine Ag grains, with the amount of oxygen vacancies growing as the calcination temperature increases. An enhanced Ag–CeO2 interfacial interaction was assumed to cause high performance of the composites in electrocatalytic oxidation of sodium borohydride. In Ref. [195] Au was shown to impair the promoting effect on these composites and decrease the reaction resistance. The activity improvement was assumed to be caused by strengthening of interfacial interaction between the Ag–Au solid solution and CeO2 particles due to Au effect, while the thermal stability and electron transport properties also improved. An increase of the Au content in the precursor alloy results in the reduction of catalytic activity and thermal stability.
In Ref. [52] 3D Ag/CeO2 nanorods with high electrocatalytic activity for NaBH4 electrooxidation were discussed. The ongoing calcination in air resulted in the dispersion of small Ag nanoparticles on the CeO2 surface, and well-defined Ag–CeO2 interfaces were created, where nanorods were connected by large conductive Ag nanoparticles. The resulting mass specific current of the composite 2.5 times exceeded the one for pure Ag in borohydride oxidation reaction. High concentration of surface oxygen species was assumed to determine the exhibited enhanced catalytic activity along with a 3D architecture of nanorod and strong metal–support interaction.
Thus, a variation of the chemical composition of Ag/CeO2 by using various promoters and modifiers allows tuning the electrocatalytic activity of the composite.

5. Conclusions and Outlook

In the present review we have summarized the recent advances and trends on the role of metal–support interaction in Ag/CeO2 composites in their catalytic performance in total oxidation of CO, soot, and VOCs. Promising photo- and electrocatalytic applications of Ag/CeO2 composites have also been discussed. The key function of the silver–ceria interaction is connected with the following major aspects:
  • the catalytic performance of Ag/CeO2 composites strongly depends on the preparation method that determines the morphology of both Ag and ceria nanoparticles, interfacial configuration and strength of metal–support interaction;
  • active surface sites are formed at the Ag–CeO2 interface, with the interfacial O atoms exhibiting different reactivity as compared to other surface O atoms, while oxygen species over Ag particles are still of importance and participate in catalysis;
  • positively charged Ag clusters facilitate the formation of surface oxygen vacancies over ceria support, while metal Ag nanoparticles promote the reduction of CeO2 nanocrystals and enhance their catalytic activity;
  • an enhanced activity of Ag/CeO2 materials is caused by the highest surface oxygen vacancy concentration, high low-temperature reducibility as well as existence of lattice oxygen species and lattice defects formed with the participation of both silver and ceria;
  • the role of impurities (such as alkali ions, carbon-containing species, etc., appeared on the surface and/or bulk of ceria during the preparation procedure and participating in transferring of electron density to O surface species) should be considered;
  • redox properties are caused by coexistence and interplay between Ag+/Ag0 and Ce3+/Ce4+ pairs;
  • high photocatalytic activity of Ag/CeO2 composites is caused by the ability of Ag nanoparticles to prolong the lifetime of photogenerated electron–hole pairs due to the effect of localized SPR and reduction of the recombination of free charges;
  • enhanced electrocatalytic activity and good electrochemical stability of Ag/CeO2 composites are connected with strong interfacial interactions between Ag and CeO2 moieties that are caused by their specific morphology and architecture, which hinder the particulate agglomeration during the long-term electrocatalytic reaction.
Thus, the configuration of the silver–ceria interface provides the enhanced catalyst performance caused by synergistic effects of silver and cerium oxide. A proper selection of preparation method allows achieving the desired features of the composites and fine-tuning the strength of electronic metal–support interactions that can be additionally improved by application of ordered supports (e.g., SBA, MCM, MOFs, etc.) and promoters. This will allow rational designing of a new generation of highly effective Ag/CeO2 composites for environmental, energy, photo- and electrocatalytic applications.

Supplementary Materials

Supplementary File 1

Acknowledgments

This research was supported by “The Tomsk State University competitiveness improvement programme”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Web Site of World Health Organization. Available online: http://www.who.int/mediacentre/factsheets/fs313/en/ (accessed on 2 May 2018).
  2. California Air Resources Board. Definitions of VOC and ROG; California Air Resources Board: Sacramento, CA, USA, 2004.
  3. Kamal, M.S.; Razzak, S.A.; Hossain, M.M. Catalytic oxidation of volatile organic compounds (VOCs)-A review. Atmos. Environ. 2016, 140, 117–134. [Google Scholar] [CrossRef]
  4. Oh, S.-H.; Hoflund, G.B. Chemical state study of palladium powder and ceria-supported palladium during low-temperature CO oxidation. J. Phys. Chem. A 2006, 110, 7609–7613. [Google Scholar] [CrossRef] [PubMed]
  5. Slavinskaya, E.M.; Gulyaev, R.V.; Zadesenets, A.V.; Stonkus, O.A.; Zaikovskii, V.I.; Shubin, Y.V.; Korenev, S.V.; Boronin, A.I. Low-temperature CO oxidation by Pd/CeO2 catalysts synthesized using the coprecipitation method. Appl. Catal. B 2015, 166–167, 91–103. [Google Scholar] [CrossRef]
  6. Hinokuma, S.; Fujii, H.; Okamoto, M.; Ikeue, K.; Machida, M. Metallic Pd nanoparticles formed by Pd–O–Ce interaction: a reason for sintering-induced activation for CO oxidation. Chem. Mater. 2010, 22, 6183–6190. [Google Scholar] [CrossRef]
  7. Jin, M.; Park, J.-N.; Shon, J.K.; Kim, J.H.; Li, Z.; Park, Y.-K.; Kim, J.M. Low temperature CO oxidation over Pd catalysts supported on highly ordered mesoporous metal oxides. Catal. Today 2012, 185, 183–190. [Google Scholar] [CrossRef]
  8. Wu, J.; Zeng, L.; Cheng, D.; Chen, F.; Zhan, X.; Gong, J. Synthesis of Pd nanoparticles supported on CeO2 nanotubes for CO oxidation at low temperatures. Chin. J. Catal. 2016, 37, 83–90. [Google Scholar] [CrossRef]
  9. Hu, Z.; Liu, X.F.; Meng, D.M.; Guo, Y.; Guo, Y.L.; Lu, G.Z. Effect of Ceria Crystal Plane on the Physicochemical and Catalytic Properties of Pd/Ceria for CO and Propane Oxidation. ACS Catal. 2016, 6, 2265–2279. [Google Scholar] [CrossRef]
  10. Si, G.; Yu, J.; Xiao, X.; Guo, X.; Huang, H.; Mao, D.; Lu, G. Boundary role of Nano-Pd catalyst supported on ceria and the approach of promoting the boundary effect. Mol. Catal. 2018, 444, 1–9. [Google Scholar] [CrossRef]
  11. Adijanto, L.; Sampath, A.; Yu, A.S.; Cargnello, M.; Fornasiero, P.; Gorte, R.J.; Vohs, J.M. Synthesis and Stability of Pd@CeO2 Core–Shell Catalyst Films in Solid Oxide Fuel Cell Anodes. ACS Catal. 2013, 3, 1801–1809. [Google Scholar] [CrossRef]
  12. Carlsson, P.A.; Skoglundh, M. Low-temperature oxidation of carbon monoxide and methane over alumina and ceria supported platinum catalysts. Appl. Catal. B. 2011, 101, 669–675. [Google Scholar] [CrossRef] [Green Version]
  13. Morfin, F.; Nguyen, T.S.; Rousset, J.L.; Piccolo, L. Synergy between hydrogen and ceria in Pt-catalyzed CO oxidation: An investigation on Pt–CeO2 catalysts synthesized by solution combustion. Appl. Catal. B 2016, 197, 2–13. [Google Scholar] [CrossRef]
  14. Lee, J.; Ryou, Y.; Kim, J.; Chan, X.; Kim, T.J.; Kim, D.H. Influence of the Defect Concentration of Ceria on the Pt Dispersion and the CO Oxidation Activity of Pt/CeO2. J. Phys. Chem. C 2018, 122, 4972–4983. [Google Scholar] [CrossRef]
  15. Peng, R.; Li, S.; Sun, X.; Ren, Q.; Chen, L.; Fu, M.; Wu, J.; Ye, D. Size effect of Pt nanoparticles on the catalytic oxidation of toluene over Pt/CeO2 catalysts. Appl. Catal. B 2018, 220, 462–470. [Google Scholar] [CrossRef]
  16. Bera, P.; Gayen, A.; Hegde, M.S.; Lalla, N.P.; Spadaro, L.; Frusteri, F.; Arena, F. Promoting Effect of CeO2 in Combustion Synthesized Pt/CeO2 Catalyst for CO Oxidation. J. Phys. Chem. B 2003, 107, 6122–6130. [Google Scholar] [CrossRef]
  17. Zhang, R.; Lu, K.; Zong, L.; Tong, S.; Wang, X.; Feng, G. Gold supported on ceria nanotubes for CO oxidation. Appl. Surf. Sci. 2017, 416, 183–190. [Google Scholar] [CrossRef]
  18. Zhang, R.; Lu, K.; Zong, L.; Tong, S.; Wang, X.; Zhou, J.; Lu, Z.-H.; Feng, G. CO Oxidation Activity at Room Temperature over Au/CeO2 Catalysts: Disclosure of Induction Period and Humidity Effect. Mol. Catal. 2017, 442, 173–180. [Google Scholar] [CrossRef]
  19. Centeno, M.A.; Reina, T.R.; Ivanova, S.; Laguna, O.H.; Odriozola, J.A. Au/CeO2 Catalysts: Structure and CO Oxidation Activity. Catalysts 2016, 6, 158. [Google Scholar] [CrossRef]
  20. El-Moemen, A.A.; Abdel-Mageed, A.M.; Bansmann, J.; Parlinska-Wojtan, M.; Behm, R.J.; Kučerová, G. Deactivation of Au/CeO2 catalysts during CO oxidation: Influence of pretreatment and reaction conditions. J. Catal. 2016, 341, 160–179. [Google Scholar] [CrossRef]
  21. Sudarsanam, P.; Mallesham, B.; Reddy, P.S.; Großmann, D.; Grünert, W.; Reddy, B.M. Nano-Au/CeO2 catalysts for CO oxidation: Influence of dopants (Fe, La and Zr) on the physicochemical properties and catalytic activity. Appl. Catal. B 2014, 144, 900–908. [Google Scholar] [CrossRef]
  22. Zhang, S.; Li, X.-S.; Chen, B.; Zhu, X.; Shi, C.; Zhu, A.-M. CO Oxidation Activity at Room Temperature over Au/CeO2 Catalysts: Disclosure of Induction Period and Humidity Effect. ACS Catal. 2014, 4, 3481–3489. [Google Scholar] [CrossRef]
  23. Li, H.-F.; Zhang, N.; Chen, P.; Luo, M.-F.; Lu, J.-Q. High surface area Au/CeO2 catalysts for low temperature formaldehyde oxidation. Appl. Catal. B 2011, 110, 279–285. [Google Scholar] [CrossRef]
  24. Satsuma, A.; Yanagihara, M.; Ohyama, J.; Shimizu, K. Oxidation of CO over Ru/ceria prepared by self-dispersion of Ru metal powder into nano-sized particle. Catal. Today 2013, 201, 62–67. [Google Scholar] [CrossRef]
  25. Vargas, E.; Simakov, A.; Rangel, R.; Castillon, F. CO oxidation over Ce–Ru–O catalysts. In Proceedings of the 20th North American Catalysis Meeting, Houston, TX, USA, 17–22 June 2007. [Google Scholar]
  26. Asadullah, M.; Fujimoto, K.; Tomishige, K. Catalytic Performance of Rh/CeO2 in the Gasification of Cellulose to Synthesis Gas at Low Temperature. Ind. Eng. Chem. Res. 2001, 40, 5894–5900. [Google Scholar] [CrossRef]
  27. Li, K.; Wang, X.; Zhou, Z.; Wu, X.; Weng, D. Oxygen Storage Capacity of Pt-, Pd-, Rh/CeO2-Based Oxide Catalyst. J. Rare Earths 2007, 25, 6–10. [Google Scholar]
  28. Kurnatowska, M.; Kepinski, L. Structure and thermal stability of nanocrystalline Ce1−xRhxO2−y in reducing and oxidizing atmosphere. Mater. Res. Bull. 2013, 48, 852–862. [Google Scholar] [CrossRef]
  29. Zhao, Y.; Teng, B.-T.; Wen, X.-D.; Zhao, Y.; Zhao, L.-H.; Luo, M.-F. A theoretical evaluation and comparison of MxCe1−xO2−δ (M = Au, Pd, Pt, and Rh) catalysts. Catal. Commun. 2012, 27, 63–68. [Google Scholar] [CrossRef]
  30. Li, Y.; Cai, Y.; Xing, X.; Chen, N.; Deng, D.; Wang, Y. Catalytic activity for CO oxidation of Cu–CeO2 composite nanoparticles synthesized by a hydrothermal method. Anal. Methods 2015, 7, 3238–3245. [Google Scholar] [CrossRef]
  31. Sundar, R.S.; Deevi, S. CO oxidation activity of Cu–CeO2 nano-composite catalysts prepared by laser vaporization and controlled condensation. J. Nanopart. Res. 2006, 8, 497–509. [Google Scholar] [CrossRef]
  32. Xu, X.; Li, J.; Hao, Z. CeO2-Co3O4 Catalysts for CO Oxidation. J. Rare Earths 2006, 24, 172–176. [Google Scholar] [CrossRef]
  33. Quiroz, J.; Giraudon, J.-M.; Gervasini, A.; Dujardin, C.; Lancelot, C.; Trentesaux, M.; Lamonier, J.-F. Total Oxidation of Formaldehyde over MnOx-CeO2 Catalysts: The Effect of Acid Treatment. ACS Catal. 2015, 5, 2260–2269. [Google Scholar] [CrossRef]
  34. Huang, H.; Xu, Y.; Feng, Q.; Leung, D.Y. Low temperature catalytic oxidation of volatile organic compounds: A review. Catal. Sci. Technol. 2015, 5, 2649–2669. [Google Scholar] [CrossRef]
  35. Vodyankina, O.V.; Blokhina, A.S.; Kurzina, I.A.; Sobolev, V.I.; Koltunov, K.Y.; Chukhlomina, L.N.; Dvilis, E.S. Selective oxidation of alcohols over Ag-containing Si3N4 catalysts. Catal. Today 2013, 203, 127–132. [Google Scholar] [CrossRef]
  36. Mamontov, G.V.; Grabchenko, M.V.; Sobolev, V.I.; Zaikovskii, V.I.; Vodyankina, O.V. Ethanol dehydrogenation over Ag-CeO2/SiO2 catalyst: Role of Ag-CeO2 interface. Appl. Catal. A 2016, 528, 161–167. [Google Scholar] [CrossRef]
  37. Dutov, V.V.; Mamontov, G.V.; Sobolev, V.I.; Vodyankina, O.V. Silica-supported silver-containing OMS-2 catalysts for ethanol oxidative dehydrogenation. Catal. Today 2016, 278, 164–173. [Google Scholar] [CrossRef]
  38. Mamontov, G.V.; Gorbunova, A.S.; Vyshegorodtseva, E.V.; Zaikovskii, V.I.; Vodyankina, O.V. Selective oxidation of CO in the presence of propylene over Ag/MCM-41 catalyst. Catal. Today 2018. [Google Scholar] [CrossRef]
  39. Pan, C.-J.; Tsai, M.-C.; Su, W.-N.; Rick, J.; Akalework, N.G.; Agegnehu, A.K.; Cheng, S.-Y.; Hwang, B.-J. Tuning/exploiting Strong Metal-Support Interaction (SMSI) in Heterogeneous Catalysis. J. Taiwan Inst. Chem. E 2017, 74, 154–186. [Google Scholar] [CrossRef]
  40. Ma, L.; Wang, D.; Li, J.; Bai, B.; Fu, L.; Li, Y. Ag/CeO2 nanospheres: Efficient catalysts for formaldehyde oxidation. Appl. Catal. B 2014, 148, 36–43. [Google Scholar] [CrossRef]
  41. Skaf, M.; Aouad, S.; Hany, S.; Cousin, R.; Abi-Aadand, E.; Aboukais, A. Physicochemical characterization and catalytic performance of 10% Ag/CeO2 catalysts prepared by impregnation and deposition-precipitation. J. Catal. 2014, 320, 137–146. [Google Scholar] [CrossRef]
  42. Qu, Z.; Yu, F.; Zhang, X.; Wang, Y.; Gao, J. Support effects on the structure and catalytic activity of mesoporous Ag/CeO2 catalysts for CO oxidation. Chem. Eng. J. 2013, 229, 522–532. [Google Scholar] [CrossRef]
  43. Chang, S.; Li, M.; Hua, Q.; Zhang, L.; Ma, Y.; Ye, B.; Huang, W. Shape-dependent interplay between oxygen vacancies and Ag–CeO2 interaction in Ag/CeO2 catalysts and their influence on the catalytic activity. J. Catal. 2012, 293, 195–204. [Google Scholar] [CrossRef]
  44. Kayama, T.; Yamazaki, K.; Shinjoh, H. Nanostructured Ceria−Silver Synthesized in a One-Pot Redox Reaction Catalyzes Carbon Oxidation. J. Am. Chem. Soc. 2010, 132, 13154–13155. [Google Scholar]
  45. Shimizu, K.; Kawachi, H.; Satsuma, A. Study of active sites and mechanism for soot oxidation by silver-loaded ceria catalyst. Appl. Catal. B Environ. 2010, 96, 169–175. [Google Scholar] [CrossRef]
  46. Aneggi, E.; Llorca, J.; de Leitenburg, C.; Dolcetti, G.; Trovarelli, A. Soot Combustion Over Silver-Supported Catalysts. Appl. Catal. B Environ. 2009, 91, 489–498. [Google Scholar] [CrossRef]
  47. Lee, C.; Park, J.; Shul, Y.-G.; Einaga, H.; Teraoka, Y. Ag supported on electrospun macro-structure CeO2 fibrous mats for diesel soot oxidation. Appl. Catal. B Environ. 2015, 174, 185–192. [Google Scholar] [CrossRef]
  48. Machida, M.; Murata, Y.; Kishikawa, K.; Zhang, D.; Ikeue, K. On the Reasons for High Activity of CeO2 Catalyst for Soot Oxidation. Chem. Mater. 2008, 20, 4489–4494. [Google Scholar] [CrossRef]
  49. Yamazaki, K.; Kayama, T.; Dong, F.; Shinjoh, H. A mechanistic study on soot oxidation over CeO2-Ag catalyst with ‘rice-ball’ morphology. J. Catal. 2011, 282, 289–298. [Google Scholar] [CrossRef]
  50. Leng, Q.; Yang, D.; Yang, Q.; Hu, C.; Kang, Y.; Wang, M.; Hashim, M. Building novel Ag/CeO2 heterostructure for enhancing photocatalytic activity. Mater. Res. Bull. 2015, 65, 266–272. [Google Scholar] [CrossRef]
  51. Li, G.; Lu, F.; Wei, X.; Song, X.; Sun, Z.; Yang, Z.; Yang, S. Nanoporous Ag–CeO2 ribbons prepared by chemical dealloying and their electrocatalytic properties. J. Mater. Chem. 2013, 1, 4974–4981. [Google Scholar] [CrossRef]
  52. Zhang, X.; Li, G.; Song, X.; Yang, S.; Sun, Z. Three-dimensional architecture of Ag/CeO2 nanorod composites prepared by dealloying and their electrocatalytic performance. RSC Adv. 2017, 7, 32442–32451. [Google Scholar] [CrossRef]
  53. Imamura, S.; Yamada, H.; Utani, K. Combustion activity of Ag/CeO2 composite catalyst. Appl. Catal. A Gen. 2000, 192, 221–226. [Google Scholar] [CrossRef]
  54. Scirè, S.; Riccobene, P.M.; Crisafulli, C. Ceria supported group IB metal catalysts for the combustion of volatile organic compounds and the preferential oxidation of CO. Appl. Catal. B Environ. 2010, 101, 109–117. [Google Scholar] [CrossRef]
  55. Fiorenza, R.; Crisafulli, C.; Condorelli, G.G.; Lupo, F.; Scirè, S. Au–Ag/CeO2 and Au–Cu/CeO2 Catalysts for Volatile Organic Compounds Oxidation and CO Preferential Oxidation. Catal. Lett. 2015, 145, 1691–1702. [Google Scholar] [CrossRef]
  56. Wang, L.; He, H.; Yu, Y.; Sun, L.; Liu, S.; Zhang, C.; He, L. Morphology-dependent bactericidal activities of Ag/CeO2 catalysts against Escherichia coli. J. Inorg. Biochem. 2014, 135, 45–53. [Google Scholar] [CrossRef] [PubMed]
  57. Beuhler, R.J.; Rao, R.M.; Hrbek, J.; White, M.G. Study of the Partial Oxidation of Methanol to Formaldehyde on a Polycrystalline Ag Foil. J. Phys. Chem. B 2001, 105, 5950–5956. [Google Scholar] [CrossRef]
  58. Mamontov, G.V.; Magaev, O.V.; Knyazev, A.S.; Vodyankina, O.V. Influence of phosphate addition on activity of Ag and Cu catalysts for partial oxidation of alcohols. Catal. Today 2013, 203, 122–126. [Google Scholar] [CrossRef]
  59. Rodriguez, J.A.; Grinter, D.C.; Liu, Z.; Palomino, R.M.; Senanayake, S.D. Ceria-based model catalysts: Fundamental studies on the importance of the metal-ceria interface in CO oxidation, the water-gas shift, CO2 hydrogenation, and methane and alcohol reforming. Chem. Soc. Rev. 2017, 46, 1824–1841. [Google Scholar] [CrossRef] [PubMed]
  60. Zagaynov, I.V.; Naumkin, A.V.; Grigoriev, Y.V. Perspective intermediate temperature ceria based catalysts for CO oxidation. Appl. Catal. B Environ. 2018, 236, 171–175. [Google Scholar] [CrossRef]
  61. Slavinskaya, E.M.; Stadnichenko, A.I.; Muravyov, V.V.; Kardash, T.Y.; Derevyannikova, E.A.; Zaikovskii, V.I.; Stonkus, O.A.; Lapin, I.N.; Svetlichnyi, V.A.; Boronin, A.I. Transformation of a Pt–CeO2 Mechanical Mixture of Pulsed-Laser-Ablated Nanoparticles to a Highly Active Catalyst for Carbon Monoxide Oxidation. ChemCatChem 2018, 10, 2232–2247. [Google Scholar] [CrossRef]
  62. Park, Y.; Kim, S.K.; Pradhan, D.; Sohn, Y. Thermal H2-treatment effects on CO/CO2 conversion over Pd-doped CeO2 comparison with Au and Ag-doped CeO2. Reac. Kinet. Mech. Cat. 2014, 113, 85–100. [Google Scholar] [CrossRef]
  63. Park, Y.; Na, Y.; Pradhan, Y.; Sohn, Y. Liquid-Phase Ethanol Oxidation and Gas-Phase CO Oxidation Reactions over M Doped (M = Ag, Au, Pd, and Ni) and MM’ Codoped CeO2 Nanoparticles. J. Catal. 2016, 2176576. [Google Scholar] [CrossRef]
  64. Dutov, V.V.; Mamontov, G.V.; Zaikovskii, V.I.; Liotta, L.F.; Vodyankina, O.V. Low-temperature CO oxidation over Ag/SiO2 catalysts: Effect of OH/Ag ratio. Appl. Catal. B 2018, 221, 598–609. [Google Scholar] [CrossRef]
  65. Dutov, V.V.; Mamontov, G.V.; Zaikovskii, V.I.; Vodyankina, O.V. The effect of support pretreatment on activity of Ag/SiO2 catalysts in low-temperature CO oxidation. Catal. Today 2016, 278, 150–156. [Google Scholar] [CrossRef]
  66. Afanasev, D.S.; Yakovina, O.A.; Kuznetsova, N.I.; Lisitsyn, A.S. High activity in CO oxidation of Ag nanoparticles supported on fumed silica. Catal. Commun. 2012, 22, 43–47. [Google Scholar] [CrossRef]
  67. Liu, H.; Ma, D.; Blackley, R.A.; Zhou, W.; Bao, X. Highly active mesostructured silica hosted silver catalysts for CO oxidation using the one-pot synthesis approach. Chem. Commun. 2008, 2677–2678. [Google Scholar] [CrossRef] [PubMed]
  68. Mamontov, G.V.; Dutov, V.V.; Sobolev, V.I.; Vodyankina, O.V. Effect of transition metal oxide additives on the activity of an Ag/SiO2 catalyst in carbon monoxide oxidation. Kinet. Catal. 2013, 54, 487–491. [Google Scholar] [CrossRef]
  69. Zhang, D.; Du, X.; Shi, L.; Gao, R. Shape-controlled synthesis and catalytic application of ceria nanomaterials. Dalton Trans. 2012, 41, 14455–14475. [Google Scholar] [CrossRef] [PubMed]
  70. Mai, H.-X.; Sun, L.-D.; Zhang, Y.-W.; Si, R.; Feng, W.; Zhang, H.-P.; Liu, H.-C.; Yan, C.-H. Shape-Selective Synthesis and Oxygen Storage Behavior of Ceria Nanopolyhedra, Nanorods, and Nanocubes. J. Phys. Chem. B. 2005, 109, 24380–24385. [Google Scholar] [CrossRef] [PubMed]
  71. Wu, Z.; Li, M.; Overbury, S.H. On the structure dependence of CO oxidation over CeO2 nanocrystals with well-defined surface planes. J. Catal. 2012, 285, 61–73. [Google Scholar] [CrossRef]
  72. Kang, Y.; Sun, M.; Li, A. Studies of the Catalytic Oxidation of CO Over Ag/CeO2 Catalyst. Catal. Lett. 2012, 142, 1498–1504. [Google Scholar] [CrossRef]
  73. Li, G.; Zhang, X.; Feng, W.; Fang, X.; Liu, J. Nanoporous CeO2–Ag catalysts prepared by etching the CeO2/CuO/Ag2O mixed oxides for CO oxidation. Corros. Sci. 2018, 134, 140–148. [Google Scholar] [CrossRef]
  74. Liang, X.; Xiao, J.; Chen, B.; Li, Y. Catalytically Stable and Active CeO2 Mesoporous Spheres. Inorg. Chem. 2010, 49, 8188–8190. [Google Scholar] [CrossRef] [PubMed]
  75. Li, G.; Tang, Z. Noble metal nanoparticle@metal oxide core/yolk–shell nanostructures as catalysts: recent progress and perspective. Nanoscale 2014, 6, 3995–4011. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, Y.; Arandiyan, H.; Scott, J.; Bagheri, A.; Dai, H.; Amal, R. Recent Advances in Ordered Meso/macroporous Metal Oxides for Heterogeneous Catalysis: A Review. J. Mater. Chem. A 2017, 5, 8825–8846. [Google Scholar] [CrossRef]
  77. Zhang, J.; Li, L.; Huang, X.; Li, G. Fabrication of Ag–CeO2 core–shell nanospheres with enhanced catalytic performance due to strengthening of the interfacial interactions. J. Mater. Chem. 2012, 22, 10480–10487. [Google Scholar] [CrossRef]
  78. Badri, A.; Binet, C.; Lavalley, J.-C. An FTIR study of surface ceria hydroxy groups during a redox process with H2. J. Chem. Soc. Faraday Trans. 1996, 92, 4669–4673. [Google Scholar] [CrossRef]
  79. Grabchenko, M.V.; Mamontov, G.V.; Zaikovskii, V.I.; La Parola., V.; Liotta, L.F.; Vodyankina, O.V. Design of Ag-CeO2/SiO2 catalyst for oxidative dehydrogenation of ethanol: Control of Ag–CeO2 interfacial interaction. Catal. Today 2018. [Google Scholar] [CrossRef]
  80. Grabchenko, M.V.; Mamontov, G.V.; Zaikovskii, V.I.; Vodyankina, O.V. Effect of the metal−support interaction in Ag/CeO2 catalysts on their activity in ethanol oxidation. Kinet. Catal. 2017, 58, 642–648. [Google Scholar] [CrossRef]
  81. Mitsudome, T.; Mikami, Y.; Matoba, M.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Design of a silver-cerium dioxide core-shell nanocomposite catalyst for chemoselective reduction reactions. Angew. Chem. Int. Ed. 2012, 51, 136–139. [Google Scholar] [CrossRef] [PubMed]
  82. Zhou, Q.; Ma, S.; Zhan, S. Superior photocatalytic disinfection effect of Ag-3D ordered mesoporous CeO2 under visible light. Appl. Catal. B 2018, 224, 27–37. [Google Scholar] [CrossRef]
  83. Stanmore, B.R.; Brilhac, J.F.; Gilot, P. The oxidation of soot: a review of experiments, mechanisms and models. Carbon 2001, 39, 2247–2268. [Google Scholar] [CrossRef]
  84. Neyertz, C.A.; Banus, E.D.; Mir’o, 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]
  85. Fino, D.; Bensaid, S.; Piumetti, M.; Russo, N. A review on the catalyticcombustion of soot in diesel particulate filters for automotive applications:from powder catalysts to structured reactors. Appl. Catal. A 2016, 509, 75–96. [Google Scholar] [CrossRef]
  86. Neeft, P.A.; Makkee, M.; Moulijn, J.A. Catalysts for the oxidation of soot from diesel exhaust gases. I. An exploratory study. Appl. Catal. B 1996, 8, 57–78. [Google Scholar] [CrossRef]
  87. Kobayashi, Y.; Hikosaka, R. Analyzing Loose Contact Oxidation of Diesel Engine Soot and Ag/CeO2 Catalyst Using Nonlinear Regression Analysis. Bull. Chem. React. Eng. Catal. 2017, 12, 14–23. [Google Scholar] [CrossRef]
  88. Piumetti, M.; Bensaid, S.; Russo, N.; Fino, D. Nanostructured ceria-based catalysts for soot combustion: Investigations on the surface sensitivity. Appl. Catal. B 2015, 165, 742–751. [Google Scholar] [CrossRef]
  89. Hernández-Giménez, A.M.; Castelló, D.L.; Bueno-López, A. Diesel soot combustion catalysts: Review of active phases. Chem. Pap. 2014, 68, 1154–1168. [Google Scholar] [CrossRef] [Green Version]
  90. Shen, Q.; Wu, M.; Wang, H.; He, C.; Hao, Z.; Wei, W.; Sun, Y. Facile synthesis of catalytically active CeO2 for soot combustion. Catal. Sci. Technol. 2015, 5, 1941–1952. [Google Scholar] [CrossRef]
  91. Gnanamani, M.K.; Jacobs, G.; Martinelli, M.; Shafer, W.D.; Hopps, S.D.; Thomas, G.A.; Davis, B.H. Dehydration of 1,5-Pentanediol over Na-Doped CeO2 Catalysts. ChemCatChem 2018, 10, 1148–1154. [Google Scholar] [CrossRef]
  92. Miceli, P.; Bensaid, S.; Russo, N.; Fino, D. CeO2-based catalysts with engineered morphologies for soot oxidation to enhance soot-catalyst contact. Nanoscale Res. Lett. 2014, 9, 254–264. [Google Scholar] [CrossRef] [PubMed]
  93. Guillén-Hurtado, N.; Bueno-López, A.; García-García, A. Catalytic performances of ceria and ceria-zirconia materials for the combustion of diesel soot under NOx/O2 and O2. Importance of the cerium precursor salt. Appl. Catal. A 2012, 437–438, 166–172. [Google Scholar] [CrossRef]
  94. Bensaid, S.; Russo, N.; Fino, D. CeO2 catalysts with fibrous morphology for soot oxidation: The importance of the soot-catalyst contact conditions. Catal. Today 2013, 216, 57–63. [Google Scholar] [CrossRef]
  95. Atribak, I.; Such-Basáñez, I.; Bueno-López, A.; García-García, A. Comparison of the catalytic activity of MO2 (M = Ti, Zr, Ce) for soot oxidation under NOx/O2. J. Catal. 2007, 250, 75–84. [Google Scholar] [CrossRef]
  96. Zhang, W.; Niu, X.; Chen, L.; Yuan, F.; Zhu, Y. Soot Combustion over Nanostructured Ceria with Different Morphologies. Sci. Rep. 2016, 6, 29062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Aneggi, E.; Wiater, D.; de Leitenburg, C.; Llorca, J.; Trovarelli, A. Shape-Dependent Activity of Ceria in Soot Combustion. ACS Catal. 2014, 4, 172–181. [Google Scholar] [CrossRef]
  98. Kaspar, J.; Fornasiero, P.; Graziani, M. Use of CeO2-based oxides in the three-way catalysis. Catal. Today 1999, 50, 285–298. [Google Scholar] [CrossRef]
  99. Mukherjee, D.; Rao, B.G.; Reddy, B.M. CO and soot oxidation activity of doped ceria: Influence of dopants. Appl. Catal. B 2016, 197, 105–115. [Google Scholar] [CrossRef]
  100. Mukherjeea, D.; Reddy, B.M. Noble metal-free CeO2-based mixed oxides for CO and soot oxidation. Catal. Today 2018, 309, 227–235. [Google Scholar] [CrossRef]
  101. Kong, D.; Wang, G.; Pan, Y.; Hu, S.; Hou, J.; Pan, H.; Campbell, C.T.; Zhu, J. Growth, Structure, and Stability of Ag on CeO2(111): Synchrotron Radiation Photoemission Studies. J. Phys. Chem. C 2011, 115, 6715–6725. [Google Scholar] [CrossRef]
  102. Liu, S.; Wu, X.; Weng, D.; . Ran, R. Ceria-based catalysts for soot oxidation: A review. J. Rare Earth 2015, 33, 567–590. [Google Scholar] [CrossRef]
  103. Severin, N.; Kirstein, S.; Sokolov, I.M.; Rabe, J.P. Rapid trench channeling of graphenes with catalytic silver nanoparticles. Nano Lett. 2009, 9, 457–461. [Google Scholar] [CrossRef] [PubMed]
  104. Yamazaki, K.; Sakakibara, Y.; Dong, F.; Shinjoh, H. The remote oxidation of soot separated by ash deposits viasilver–ceria composite catalysts. Appl. Catal. A 2014, 476, 113–120. [Google Scholar] [CrossRef]
  105. Liu, S.; Wu, X.; Liu, W.; Chen, W.; Ran, R.; Li, M.; Weng, D. Soot oxidation over CeO2 and Ag/CeO2: Factors determining the catalyst activity and stability during reaction. J. Catal. 2016, 337, 188–198. [Google Scholar] [CrossRef]
  106. Wu, S.; Yang, Y.; Lu, C.; Ma, Y.; Yuan, S.; Qian, G. Soot oxidation over CeO2 or Ag/CeO2: influences of bulk oxygen vacancies and surface oxygen vacancies on activity and stability of catalyst. Eur. J. Inorg. Chem. 2018, 2018, 2944–2951. [Google Scholar] [CrossRef]
  107. Piumetti, M.; van der Linden, B.; Makkee, M.; Miceli, P.; Fino, D.; Russo, N.; Bensaid, S. Contact dynamics for a solid–solid reaction mediated by gas-phase oxygen: Study on the soot oxidation over ceria-based catalysts. Appl. Catal. B 2016, 199, 96–107. [Google Scholar] [CrossRef]
  108. Shangguan, W.F.; Teraoka, Y.; Kagawa, S. Kinetics of soot–O2, soot–NO and soot–O2–NO reactions over spinel-type CuFe2O4 catalyst. Appl. Catal. B 1997, 12, 237–247. [Google Scholar] [CrossRef]
  109. Aneggi, E.; Leitenburg, C.; Trovarelli, A. On the role of lattice/surface oxygen in ceria-zirconia catalysts for diesel soot combustion. Catal. Today 2012, 181, 108–115. [Google Scholar] [CrossRef]
  110. Bueno-López, A. Diesel soot combustion ceria catalysts. Appl. Catal. B 2014, 146, 1–11. [Google Scholar] [CrossRef] [Green Version]
  111. Bueno-López, A.; Krishna, K.; Makkee, M.; Moulijn, J.A. Enhanced soot oxidation by lattice oxygen via La3+–doped CeO2. J. Catal. 2005, 230, 237–248. [Google Scholar] [CrossRef]
  112. Wang, H.; Liu, S.; Zhao, Z.; Zou, X.; Liu, M.; Liu, W.; Wu, X.; Weng, D. Activation and deactivation of Ag/CeO2 during soot oxidation: influences of interfacial ceria reduction. Catal. Sci. Technol. 2017, 7, 2129–2139. [Google Scholar] [CrossRef]
  113. Hosokawa, S.; Taniguchi, M.; Utani, K.; Kanai, H.; Imamura, S. Affinity order among noble metals and CeO2. Appl. Catal. A. 2005, 289, 115–120. [Google Scholar] [CrossRef]
  114. Nagai, Y.; Hirabayashi, T.; Dohmae, K.; Takagi, N.; Minami, T.; Shinjoh, H.; Matsumoto, S. Sintering inhibition mechanism of platinum supported on ceria-based oxide and Ptoxide–support interaction. J. Catal. 2006, 242, 103–109. [Google Scholar] [CrossRef]
  115. Nagai, Y.; Dohmae, K.; Ikeda, Y.; Takagi, N.; Hara, N.; Tanabe, T.; Guilera, G.; Pascarelli, S.; Newton, M.A.; Takahashi, N.; et al. In situ observation of platinum sintering on ceria-based oxide for autoexhaust catalysts using Turbo-XAS. Catal. Today 2011, 175, 133–140. [Google Scholar] [CrossRef]
  116. Malhautier, L.; Quijano, G.; Avezac, M.; Rocher, J.; Fanlo, J.L. Kinetic characterization of toluene biodegradation by Rhodococcus erythropolis: towards a rationale for microflora enhancement in bioreactors devoted to air treatment. Chem. Eng. J. 2014, 247, 199–204. [Google Scholar] [CrossRef]
  117. Li, L.; Liu, S.; Liu, J. Surface modification of coconut shell based activated carbon for the improvement of hydrophobic VOC removal. J. Hazard. Mater. 2011, 192, 683–690. [Google Scholar] [CrossRef] [PubMed]
  118. Thévenet, F.; Sivachandiran, L.; Guaitella, O.; Barakat, C.; Rousseau, A. Plasma–catalyst coupling for volatile organic compound removal and indoor air treatment: a review. J. Phys. D Appl. Phys. 2014, 47, 224011. [Google Scholar] [CrossRef]
  119. Destaillats, H.; Sleiman, M.; Sullivan, D.P.; Jacquiod, C.; Sablayrolles, J.; Molins, L. Key parameters influencing the performance of photocatalytic oxidation (PCO) air purification under realistic indoor conditions. Appl. Catal. B 2012, 128, 159–170. [Google Scholar] [CrossRef]
  120. Yuan, M.H.; Chang, C.Y.; Shie, J.L.; Chang, C.C.; Chen, J.H.; Tsai, W.T. Destruction of naphthalene via ozone-catalytic oxidation process over Pt/Al2O3 catalyst. J. Hazard. Mater. 2010, 175, 809–815. [Google Scholar] [CrossRef] [PubMed]
  121. Liotta, L.F. Catalytic oxidation of volatile organic compounds on supported noble metals. Appl. Catal. B 2010, 100, 403–412. [Google Scholar] [CrossRef]
  122. Everaert, K.; Baeyens, J. Catalytic combustion of volatile organic compounds. J. Hazard. Mater. 2004, 109, 113–139. [Google Scholar] [CrossRef] [PubMed]
  123. Armor, J. N. Environmental catalysis. Appl. Catal. B 1992, 1, 221–256. [Google Scholar] [CrossRef]
  124. Spivey, J.J. Complete catalytic oxidation of volatile organics. Ind. Eng. Chem. Res. 1987, 26, 2165–2180. [Google Scholar] [CrossRef]
  125. Ordóñez, S.; Bello, L.; Sastre, H.; Rosal, R.; Dıez, F.V. Kinetics of the deep oxidation of benzene, toluene, n-hexane and their binary mixtures over a platinum on γ-alumina catalyst. Appl. Catal. B 2002, 38, 139–149. [Google Scholar] [CrossRef]
  126. Huang, H.; Hu, P.; Huang, H.; Chen, J.; Ye, X.; Leung, D.Y. Highly dispersed and active supported Pt nanoparticles for gaseous formaldehyde oxidation: Influence of particle size. Chem. Eng. J. 2014, 252, 320–326. [Google Scholar] [CrossRef]
  127. Huang, S.; Zhang, C.; He, H. Complete oxidation of o-xylene over Pd/Al2O3 catalyst at low temperature. Catal. Today 2008, 139, 15–23. [Google Scholar] [CrossRef]
  128. Qi, J.; Chen, J.; Li, G.; Li, S.; Gao, Y.; Tang, Z. Facile synthesis of core–shell Au@ CeO2 nanocomposites with remarkably enhanced catalytic activity for CO oxidation. Energy Environ. Sci. 2012, 5, 8937–8941. [Google Scholar] [CrossRef]
  129. Hosseini, M.; Barakat, T.; Cousin, R.; Aboukaïs, A.; Su, B.L.; De Weireld, G.; Siffert, S. Catalytic performance of core–shell and alloy Pd–Au nanoparticles for total oxidation of VOC: the effect of metal deposition. Appl. Catal. B 2012, 111, 218–224. [Google Scholar] [CrossRef]
  130. Xu, R.; Wang, X.; Wang, D.; Zhou, K.; Li, Y. Surface structure effects in nanocrystal MnO2 and Ag/MnO2 catalytic oxidation of CO. J. Catal. 2006, 237, 426–430. [Google Scholar] [CrossRef]
  131. Bai, B.; Arandiyan, H.; Li, J. Comparison of the performance for oxidation of formaldehyde on nano-Co3O4, 2D-Co3O4, and 3D-Co3O4 catalysts. Appl. Catal. B 2013, 142, 677–683. [Google Scholar] [CrossRef]
  132. Montini, T.; Melchionna, M.; Monai, M.; Fornasiero, P. Fundamentals and catalytic applications of CeO2–based materials. Chem. Rev. 2016, 116, 5987–6041. [Google Scholar] [CrossRef] [PubMed]
  133. Hanafiah, M.A.K.M.; Hussin, Z.M.; Ariff, N.F.M.; Ngah, W.S.W.; Ibrahim, S.C. Monosodium glutamate functionalized chitosan beads for adsorption of precious cerium ion. Adv. Mater. Res. 2014, 970, 198–203. [Google Scholar] [CrossRef]
  134. Min, C.K. Nanostructured Pt/MnO2 Catalysts and Their Performance for Oxygen Reduction Reaction in Air Cathode Microbial Fuel Cell. Ph.D. Thesis, University Malaysia Pahang, Pekan, Pahang, Malaysia, June 2014. [Google Scholar]
  135. Abdel-Mageed, A.M.; Kučerová, G.; El-Moemen, A.A.; Bansmann, J.; Widmann, D.; Behm, R.J. Geometric and electronic structure of Au on Au/CeO2 catalysts during the CO oxidation: Deactivation by reaction induced particle growth. J. Phys. Conf. Ser. 2016, 712, 012044. [Google Scholar] [CrossRef]
  136. Tan, H.; Wang, J.; Yu, S.; Zhou, K. Support morphology-dependent catalytic activity of Pd/CeO2 for formaldehyde oxidation. Environ. Sci. Technol. 2015, 49, 8675–8682. [Google Scholar] [CrossRef] [PubMed]
  137. Zhang, J.; Li, Y.; Zhang, Y.; Chen, M.; Wang, L.; Zhang, C.; He, H. Effect of support on the activity of Ag-based catalysts for formaldehyde oxidation. Sci. Rep. 2015, 5, 12950. [Google Scholar] [CrossRef] [PubMed]
  138. Li, G.; Li, L. Highly efficient formaldehyde elimination over meso-structured M/CeO2 (M=Pd, Pt, Au and Ag) catalyst under ambient conditions. RSC Adv. 2015, 5, 36428–36433. [Google Scholar] [CrossRef]
  139. Yu, L.; Peng, R.; Chen, L.; Fu, M.; Wu, J.; Ye, D. Ag supported on CeO2 with different morphologies for the catalytic oxidation of HCHO. Chem. Eng. J. 2018, 334, 2480–2487. [Google Scholar] [CrossRef]
  140. Imamura, S.; Uchihori, D.; Utani, K.; Ito, T. Oxidative decomposition of formaldehyde on silver-cerium composite oxide catalyst. Catal. Lett. 1994, 24, 377–384. [Google Scholar] [CrossRef]
  141. Benaissa, S.; Chérif-Aouali, L.; Siffert, S.; Aboukais, A.; Cousin, R.; Bengueddach, A. New Nanosilver/Ceria Catalyst for Atmospheric Pollution Treatment. Nano 2015, 10, 1550043. [Google Scholar] [CrossRef]
  142. Aboukais, A.; Skaf, M.; Hany, S.; Cousin, R.; Aouad, S.; Labaki, M.; Abi-Aad, E. A comparative study of Cu, Ag and Au doped CeO2 in the total oxidation of volatile organic compounds (VOCs). Mater. Chem. Phys. 2016, 177, 570–576. [Google Scholar] [CrossRef]
  143. Liu, M.; Wu, X.; Liu, S.; Gao, Y.; Chen, Z.; Ma, Y.; Weng, D. Study of Ag/CeO2 catalysts for naphthalene oxidation: Balancing the oxygen availability and oxygen regeneration capacity. Appl. Catal. B 2017, 219, 231–240. [Google Scholar] [CrossRef]
  144. Yang, H.; Deng, J.; Liu, Y.; Xie, S.; Wu, Z.; Dai, H. Preparation and catalytic performance of Ag, Au, Pd or Pt nanoparticles supported on 3DOM CeO2–Al2O3 for toluene oxidation. J. Mol. Catal. A Chem. 2016, 414, 9–18. [Google Scholar] [CrossRef]
  145. Kharlamova, T.; Mamontov, G.; Salaev, M.; Zaikovskii, V.; Popova, G.; Sobolev, V.; Vodyankina, O. Silica-supported silver catalysts modified by cerium/manganese oxides for total oxidation of formaldehyde. Appl. Catal. A 2013, 467, 519–529. [Google Scholar] [CrossRef]
  146. Mullins, D.R. The surface chemistry of cerium oxide. Surf. Sci. Rep. 2015, 70, 42–85. [Google Scholar] [CrossRef] [Green Version]
  147. Nolan, N. Surface Effects in the Reactivity of Ceria: A First Principles Perspective. Catal. Mater. Well-Defined Struct. 2015, 159–192. [Google Scholar]
  148. Plata, J.J.; Márquez, A.M.; Fdez Sanz, J. Improving the density functional theory+U description of CeO2 by including the contribution of the O 2p electrons. J. Chem. Phys. 2012, 136, 041101. [Google Scholar] [CrossRef] [PubMed]
  149. Kozlov, S.M.; Viñes, F.; Nilius, N.; Shaikhutdinov, S.; Neyman, K.M. Absolute surface step energies: Accurate theoretical methods applied to ceria nanoislands. J. Phys. Chem. Lett. 2012, 3, 1956–1961. [Google Scholar] [CrossRef]
  150. Nolan, M. Hybrid density functional theory description of oxygen vacancies in the CeO2 (110) and (100) surfaces. Chem. Phys. Lett. 2010, 499, 126–130. [Google Scholar] [CrossRef]
  151. Paier, J.; Penschke, C.; Sauer, J. Oxygen defects and surface chemistry of ceria: Quantum chemical studies compared to experiment. Chem. Rev. 2013, 113, 3949–3985. [Google Scholar] [CrossRef] [PubMed]
  152. Preda, G.; Pacchioni, G. Formation of oxygen active species in Ag-modified CeO2 catalyst for soot oxidation: A DFT study. Catal. Today 2011, 177, 31–38. [Google Scholar] [CrossRef]
  153. Zhu, K.-J.; Liu, J.; Yang, Y.-J.; Xu, Y.-X.; Teng, B.-T.; Wen, X.-D.; Fan, M. A method to explore the quantitative interactions between metal and ceria for M/CeO2 catalysts. Surf. Sci. 2018, 669, 79–86. [Google Scholar] [CrossRef]
  154. Tereshchuk, P.; Freire, R.L.H.; Ungureanu, C.G.; Seminovski, Y.; Kiejnac, A.; Da Silva, J.L.F. The role of charge transfer in the oxidation state change of Ce atoms in the TM13–CeO2(111) systems (TM = Pd, Ag, Pt, Au): A DFT + U investigation. Phys. Chem. Chem. Phys. 2015, 17, 13520–13530. [Google Scholar] [CrossRef] [PubMed]
  155. Piotrowski, M.J.; Tereshchuk, P.; Da Silva, J.L.F. Theoretical Investigation of Small Transition-Metal Clusters Supported on the CeO2 (111) Surface. J. Phys. Chem. C 2014, 118, 21438–21446. [Google Scholar] [CrossRef]
  156. Chen, L.-J.; Tang, Y.; Cui, L.; Ouyang, C.; Shi, S. Charge transfer and formation of Ce3+ upon adsorption of metal atom M (M = Cu, Ag, Au) on CeO2 (100) surface. J. Power Sources 2013, 234, 69–81. [Google Scholar] [CrossRef]
  157. Luches, P.; Pagliuca, F.; Valeri, S.; Illas, F.; Preda, G.; Pacchioni, G. Nature of Ag Islands and Nanoparticles on the CeO2 (111) Surface. J. Phys. Chem. C 2012, 116, 1122–1132. [Google Scholar] [CrossRef]
  158. Cui, L.; Tang, Y.; Zhang, H.; Hector, L.G., Jr.; Ouyang, C.; Shi, S.; Lib, H.; Chen, L. First-principles investigation of transition metal atom M (M = Cu, Ag, Au) adsorption on CeO2 (110). Phys. Chem. Chem. Phys. 2012, 14, 1923–1933. [Google Scholar] [CrossRef] [PubMed]
  159. Tang, Y.; Zhang, H.; Cui, L.; Ouyang, C.; Shi, S.; Tang, W.; Li, H.; Chen, L. Electronic states of metal (Cu, Ag, Au) atom on CeO2 (111) surface: The role of local structural distortion. J. Power Sources 2012, 197, 28–37. [Google Scholar]
  160. Branda, M.M.; Hernández, N.C.; Sanz, J.F.; Illas, F. Density functional theory study of the interaction of Cu, Ag, and Au atoms with the regular CeO2 (111) surface. J. Phys. Chem. C 2010, 114, 1934–1941. [Google Scholar] [CrossRef]
  161. Hay, P.J.; Martin, R.L.; Uddin, J.; Scuseria, G.E. Theoretical study of CeO2 and Ce2O3 using a screened hybrid density functional. J. Chem. Phys. 2006, 125, 034712. [Google Scholar] [CrossRef] [PubMed]
  162. Da Silva, J.L.F.; Ganduglia-Pirovano, M.V.; Sauer, J.; Bayer, V.; Kresse, G. Hybrid functionals applied to rare-earth oxides: The example of ceria. Phys. Rev. B 2007, 75, 045121. [Google Scholar] [CrossRef]
  163. Yang, Z.; Yu, X.; Lu, Z.; Li, S.; Hermansson, K. Oxygen vacancy pairs on CeO2 (110): A DFT + U study. Phys. Lett. A 2009, 373, 2786–2792. [Google Scholar] [CrossRef]
  164. Dudarev, S.L.; Botton, G.A.; Savrasov, S.Y.; Humphreys, C.J.; Sutton, A.P. Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+U study. Phys. Rev. B 1998, 57, 1505–1509. [Google Scholar] [CrossRef]
  165. Branda, M.M.; Castellani, N.J.; Grau-Crespo, R.; de Leeuw, N.H.; Hernandez, N.C.; Sanz, J.F.; Neyman, K.M.; Illas, F.J. On the difficulties of present theoretical models to predict the oxidation state of atomic Au adsorbed on regular sites of CeO2(111). J. Chem. Phys. 2009, 131, 094702. [Google Scholar]
  166. Trovarelli, A.; Fornasiero, P. Catalysis by Ceria and Related Materials, 2nd ed.; Imperial College Press: London, UK, 2013. [Google Scholar]
  167. Wen, X.-J.; Niu, C.-G.; Zhang, L.; Liang, C.; Zeng, G.-M. A novel Ag2O/CeO2 heterojunction photocatalysts for photocatalytic degradation of enrofloxacin: possible degradation pathways, mineralization activity and an in depth mechanism insight. Appl. Catal. B 2018, 221, 701–714. [Google Scholar] [CrossRef]
  168. Xie, S.; Wang, Z.; Cheng, F.; Zhang, P.; Mai, W.; Tong, Y. Ceria and ceria-based nanostructured materials for photoenergy applications. Nano Energy 2017, 34, 313–337. [Google Scholar] [CrossRef]
  169. Channei, D.; Inceesungvorn, B.; Wetchakun, N.; Ukritnukun, S.; Nattestad, A.; Chen, J.; Phanichphant, S. Photocatalytic degradation of methyl orange by CeO2 and Fe–doped CeO2 films under visible light irradiation. Sci. Rep. 2014, 4, 5757. [Google Scholar] [CrossRef] [PubMed]
  170. Ren, H.; Koshy, P.; Chen, W.-F.; Qi, S.; Sorrell, C.C. Photocatalytic materials and technologies for air purification. J. Hazard. Mater. 2017, 325, 340–366. [Google Scholar] [CrossRef] [PubMed]
  171. Li, Y.; Sun, Q.; Kong, M.; Shi, W.; Huang, J.; Tang, J.; Zhao, X. Coupling oxygen ion conduction to photocatalysis in mesoporous nanorod-like ceria significantly improves photocatalytic efficiency. J. Phys. Chem. C 2011, 115, 14050–14057. [Google Scholar] [CrossRef]
  172. Sabari Arul, N.; Mangalaraj, D.; Whan Kim, T. Photocatalytic degradation mechanisms of self-assembled rose-flower-like CeO2 hierarchical nanostructures. Appl. Phys. Lett. 2013, 102, 223115. [Google Scholar] [CrossRef]
  173. Ko, J.W.; Kim, J.H.; Park, C.B. Synthesis of visible light-active CeO2 sheets via mussel-inspired CaCO3 mineralization. J. Mater. Chem. A 2013, 1, 241–245. [Google Scholar] [CrossRef]
  174. Huang, Y.; Long, B.; Tang, M.; Rui, Z.; Balogun, M.-S.; Tong, Y.; Ji, H. Bifunctional catalytic material: An ultrastable and high-performance surface defect CeO2 nanosheets for formaldehyde thermal oxidation and photocatalytic oxidation. Appl. Catal. B 2016, 181, 779–787. [Google Scholar] [CrossRef]
  175. Xu, B.; Zhang, Q.; Yuan, S.; Zhang, M.; Ohno, T. Morphology control and photocatalytic characterization of yttrium-doped hedgehog-like CeO2. Appl. Catal. B 2015, 164, 120–127. [Google Scholar] [CrossRef]
  176. Ji, Y.; Ferronato, C.; Salvador, A.; Yang, X.; Chovelon, J.M. Degradation of ciprofloxacin and sulfamethoxazole by ferrous-activated persulfate: implications for remediation of groundwater contaminated by antibiotics. Sci. Total Environ. 2014, 472, 800–808. [Google Scholar] [CrossRef] [PubMed]
  177. Huang, G.F.; Ma, Z.L.; Huang, W.Q.; Tian, Y.; Jiao, C.; Yang, Z.M.; Wan, Z.; Pan, A. Ag3PO4 Semiconductor Photocatalyst: Possibilities and Challenges. J. Nanomater. 2013, 8, 371356. [Google Scholar]
  178. Yang, Z.-M.; Huang, G.-F.; Huang, W.-Q.; Wei, J.-M.; Yan, X.-G.; Liu, Y.-Y.; Jiao, C.; Wan, Z.; Pan, A. Novel Ag3PO4/CeO2 composite with high efficiency and stability for photocatalytic applications. J. Mater. Chem. A 2014, 2, 1750–1756. [Google Scholar] [CrossRef]
  179. Wu, C. Synthesis of Ag2CO3/CeO2 microcomposite with visible light-driven photocatalytic activity. Mater. Lett. 2015, 152, 76–78. [Google Scholar] [CrossRef]
  180. Wen, X.-J.; Niu, C.-G.; Huang, D.-W.; Zhang, L.; Liang, C.; Zeng, G.-M. Study of the photocatalytic degradation pathway of norfloxacin and mineralization activity using a novel ternary Ag/AgCl-CeO2 photocatalyst. J. Catal. 2017, 355, 73–86. [Google Scholar] [CrossRef]
  181. Linic, S.; Christopher, P.; Ingram, D.B. Plasmonic-metal nanostructures for efficient conversion of solar to chemical energy. Nat. Mater. 2011, 10, 911. [Google Scholar] [CrossRef] [PubMed]
  182. Tanaka, A.; Hashimoto, K.; Kominami, H. Gold and copper nanoparticles supported on cerium(IV) oxide a photocatalyst mineralizing organic acids under red light irradiation. ChemCatChem 2011, 3, 1619–1623. [Google Scholar] [CrossRef]
  183. Kominami, H.; Tanaka, A.; Hashimoto, K. Mineralization of organic acids in aqueous suspensions of gold nanoparticles supported on cerium(IV) oxide powder under visible light irradiation. Chem. Commun. 2010, 46, 1287–1289. [Google Scholar] [CrossRef] [PubMed]
  184. Liu, I.-T.; Hon, M.-H.; Kuan, C.-Y.; Teoh, L.-G. Structure and optical properties of Ag/CeO2 nanocomposites. Appl. Phys. A 2013, 111, 1181–1186. [Google Scholar] [CrossRef]
  185. Liu, T.; Li, B.; Wang, Y.; Ge, Z.; Shi, J. Facile Synthesis of Ag/CeO2 Mesoporous Composites with Enhanced Visible Light Photocatalytic Properties. Asian J. Chem. 2014, 26, 1355–1357. [Google Scholar]
  186. Hao, Y.; Li, L.; Liu, D.; Yu, H.; Zhou, Q. The synergy of SPR effect and Z-scheme of Ag on enhanced photocatalytic performance of 3DOM Ag/CeO2-ZrO2 composite. Mol. Catal. 2018, 447, 37–46. [Google Scholar] [CrossRef]
  187. Saravanan, R.; Agarwal, S.; Gupta, V.K.; Khan, M.M.; Gracia, F.; Mosquera, E.; Narayanan, V.; Stephen, A. Line defect Ce3+ induced Ag/CeO2/ZnO nanostructure for visible-light photocatalytic activity. J. Photochem. Photobiol. A 2018, 353, 499–506. [Google Scholar] [CrossRef]
  188. Mittal, M.; Gupta, A.; Pandey, O.P. Role of oxygen vacancies in Ag/Au doped CeO2 nanoparticles for fast photocatalysis. Sol. Energy 2018, 165, 206–216. [Google Scholar] [CrossRef]
  189. Barsuk, D.; Zadick, A.; Chatenet, M.; Georgarakis, K.; Panagiotopoulos, N.T.; Champion, Y.; Moreira Jorge, A., Jr. Nanoporous silver for electrocatalysis application in alkaline fuel cells. Mater. Des. 2016, 111, 528–536. [Google Scholar] [CrossRef]
  190. Lu, Q.; Rosen, J.; Zhou, Y.; Hutchings, G.S.; Kimmel, Y.C.; Chen, J.G.; Jiao, F. A selective and efficient electrocatalyst for carbon dioxide reduction. Nat. Commun. 2014, 5, 3242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Gonzalez-Macia, L.; Smyth, M.R.; Killard, A.J. Evaluation of a silver-based electrocatalyst for the determination of hydrogen peroxide formed via enzymatic oxidation. Talanta 2012, 99, 989–996. [Google Scholar] [CrossRef] [PubMed]
  192. Sreeremya, T.S.; Krishnan, A.; Remani, K.C.; Patil, K.R.; Brougham, D.F.; Ghosh, S. Shape-selective oriented cerium oxide nanocrystals permit assessment of the effect of the exposed facets on catalytic activity and oxygen storage capacity. ACS Appl. Mater. Interfaces 2015, 7, 8545–8555. [Google Scholar] [CrossRef] [PubMed]
  193. Meher, S.K.; Rao, G.R. Novel nanostructured CeO2 as efficient catalyst for energy and environmental applications. J. Chem. Sci. 2014, 126, 361–372. [Google Scholar] [CrossRef]
  194. Melchionna, M.; Fornasiero, P. The role of ceria-based nanostructured materials in energy applications. Mater. Today 2014, 17, 349–357. [Google Scholar] [CrossRef]
  195. Li, G.; Zhang, X.; Wang, L.; Song, X.; Sun, Z. Promoting Effect of Au on the Nanoporous Ag/CeO2 Composites Prepared by Dealloying for Borohydride Electro-Oxidation. J. Electrochem. Soc. 2013, 160, 1116–1122. [Google Scholar] [CrossRef]
  196. Sun, S.; Xue, Y.; Wang, Q.; Li, S.; Huang, H.; Miao, H.; Liu, Z. Electrocatalytic activity of silver decorated ceria microspheres for the oxygen reduction reaction and their application in aluminium–air batteries. Chem. Commun. 2017, 53, 7921–7924. [Google Scholar] [CrossRef] [PubMed]
Figure 1. TEM (a) and HRTEM (b) images of CeO2 nanopolyhedra. TEM (c) and HRTEM (d) images of CeO2 nanorods, inset is a fast Fourier transform (FFT) analysis. TEM (e) and HRTEM (f) images of CeO2 nanocubes, inset is a FFT analysis. Reproduced from Ref. [70] with the permission from ACS Publications.
Figure 1. TEM (a) and HRTEM (b) images of CeO2 nanopolyhedra. TEM (c) and HRTEM (d) images of CeO2 nanorods, inset is a fast Fourier transform (FFT) analysis. TEM (e) and HRTEM (f) images of CeO2 nanocubes, inset is a FFT analysis. Reproduced from Ref. [70] with the permission from ACS Publications.
Catalysts 08 00285 g001
Figure 2. Raman spectra of different catalysts under different reaction conditions (a) Ag/CeO2—5 vol% CO/N2, (b) CeO2—5 vol% CO/N2, (c) Ag/CeO2–O2, (d) CeO2–O2. Reproduced from Ref. [72] with the permission from Springer.
Figure 2. Raman spectra of different catalysts under different reaction conditions (a) Ag/CeO2—5 vol% CO/N2, (b) CeO2—5 vol% CO/N2, (c) Ag/CeO2–O2, (d) CeO2–O2. Reproduced from Ref. [72] with the permission from Springer.
Catalysts 08 00285 g002
Figure 3. (a and b) TEM images with different magnifications of CeO2 mesoporous spheres supported by a Ag nanoparticle catalyst. (c) Darkfield scanning TEM image of a single CeO2 mesoporous sphere. (d) Compositional line profile across the single sphere (from A to B) probed by Energy Dispersive X-ray Analysis (EDXA) line scanning. Reproduced from Ref. [74] with the permission from the ACS Publisher.
Figure 3. (a and b) TEM images with different magnifications of CeO2 mesoporous spheres supported by a Ag nanoparticle catalyst. (c) Darkfield scanning TEM image of a single CeO2 mesoporous sphere. (d) Compositional line profile across the single sphere (from A to B) probed by Energy Dispersive X-ray Analysis (EDXA) line scanning. Reproduced from Ref. [74] with the permission from the ACS Publisher.
Catalysts 08 00285 g003
Figure 4. FESEM images representing a loose contact mixture of CeO2 SA-stars and soot at × 40,000 (a) × 150,000 (b) level of magnifications. Reproduced from Ref. [92] with the permission from the Royal society of chemistry.
Figure 4. FESEM images representing a loose contact mixture of CeO2 SA-stars and soot at × 40,000 (a) × 150,000 (b) level of magnifications. Reproduced from Ref. [92] with the permission from the Royal society of chemistry.
Catalysts 08 00285 g004
Figure 5. Total soot conversion in loose contact conditions. Reproduced from Ref. [92] with the permission from the Royal society of chemistry.
Figure 5. Total soot conversion in loose contact conditions. Reproduced from Ref. [92] with the permission from the Royal society of chemistry.
Catalysts 08 00285 g005
Figure 6. Effect of Ag loading on soot combustion profiles of CeO2. Soot/CeO2 tight-contact mixtures with a weight ratio of 1/20 were heated in 10% O2/N2 at the rate of 10 °C·min−1. Reproduced from Ref. [48] with the permission from the American chemical society.
Figure 6. Effect of Ag loading on soot combustion profiles of CeO2. Soot/CeO2 tight-contact mixtures with a weight ratio of 1/20 were heated in 10% O2/N2 at the rate of 10 °C·min−1. Reproduced from Ref. [48] with the permission from the American chemical society.
Catalysts 08 00285 g006
Figure 7. Soot oxidation activity (Ti) of metal-loaded CeO2 measured in a flow of 10% O2 and N2 balance. Tight-contact soot/catalyst mixtures with a weight ratio of 1/20 were heated at the rate of 10 °C·min−1. Reproduced from Ref. [48] with the permission from the American chemical society.
Figure 7. Soot oxidation activity (Ti) of metal-loaded CeO2 measured in a flow of 10% O2 and N2 balance. Tight-contact soot/catalyst mixtures with a weight ratio of 1/20 were heated at the rate of 10 °C·min−1. Reproduced from Ref. [48] with the permission from the American chemical society.
Catalysts 08 00285 g007
Figure 8. Schematic illustration of Ag/CeO2 nanofiber synthesis sequence. CeO2 nanofibers were fabricated through the electrospinning of spinnable Ce/PVP in a DMF/EtOH precursor solution followed by thermal treatment. Ag was then loaded on the surfaces of the CeO2 nanofibers. Reproduced from Ref. [47] with the permission from the Elsevier.
Figure 8. Schematic illustration of Ag/CeO2 nanofiber synthesis sequence. CeO2 nanofibers were fabricated through the electrospinning of spinnable Ce/PVP in a DMF/EtOH precursor solution followed by thermal treatment. Ag was then loaded on the surfaces of the CeO2 nanofibers. Reproduced from Ref. [47] with the permission from the Elsevier.
Catalysts 08 00285 g008
Figure 9. TG and DTG curves: (a) CeO2-500, (b) CeO2-800, (c) CeO2-1000, (d) Ag/CeO2-500, (e) Ag/CeO2-800, and (f) Ag/CeO2-1000. Reproduced from Ref. [47] with the permission from the Elsevier.
Figure 9. TG and DTG curves: (a) CeO2-500, (b) CeO2-800, (c) CeO2-1000, (d) Ag/CeO2-500, (e) Ag/CeO2-800, and (f) Ag/CeO2-1000. Reproduced from Ref. [47] with the permission from the Elsevier.
Catalysts 08 00285 g009
Figure 10. A schematic mechanism of soot oxidation over Ag/CeO2 catalyst. Reproduced from Ref. [45] with the permission from the Elsevier.
Figure 10. A schematic mechanism of soot oxidation over Ag/CeO2 catalyst. Reproduced from Ref. [45] with the permission from the Elsevier.
Catalysts 08 00285 g010
Figure 11. A schematic mechanism for soot oxidation over the CeO2–Ag catalyst. Reproduced from Ref. [49] with the permission from the Elsevier.
Figure 11. A schematic mechanism for soot oxidation over the CeO2–Ag catalyst. Reproduced from Ref. [49] with the permission from the Elsevier.
Catalysts 08 00285 g011
Figure 12. A schematic mechanism for remote catalytic soot oxidation over a catalyst composed of Ag and CeO2. Reproduced from Ref. [104] with the permission from the Elsevier.
Figure 12. A schematic mechanism for remote catalytic soot oxidation over a catalyst composed of Ag and CeO2. Reproduced from Ref. [104] with the permission from the Elsevier.
Catalysts 08 00285 g012
Figure 13. Schematic explanation for activity variations over (a) Ag/CeO2 with low initial VO-s concentration (e.g., AgCe-0, AgCe-0.01, AgCe-0.02 and AgCe-0.03) and (b) Ag/CeO2 with high initial VO-s concentration (e.g., AgCe-0.04 and AgCe-0.05) during isothermal soot oxidation. Reproduced from Ref. [112] with the permission from the Royal Society of Chemistry.
Figure 13. Schematic explanation for activity variations over (a) Ag/CeO2 with low initial VO-s concentration (e.g., AgCe-0, AgCe-0.01, AgCe-0.02 and AgCe-0.03) and (b) Ag/CeO2 with high initial VO-s concentration (e.g., AgCe-0.04 and AgCe-0.05) during isothermal soot oxidation. Reproduced from Ref. [112] with the permission from the Royal Society of Chemistry.
Catalysts 08 00285 g013
Figure 14. (A) EPR spectra for (a) CeO2, (b) 10% Ag/CeO2 (Imp), and (c) 10% Ag/CeO2 (DP) recorded at −196 °C after treatment under vacuum for 30 min. (B) Propylene conversion over (a) CeO2, (b) 10% Ag/CeO2 (Imp), and (c) 10% Ag/CeO2 (DP). Reprinted from [41] with the permission of the Elsevier.
Figure 14. (A) EPR spectra for (a) CeO2, (b) 10% Ag/CeO2 (Imp), and (c) 10% Ag/CeO2 (DP) recorded at −196 °C after treatment under vacuum for 30 min. (B) Propylene conversion over (a) CeO2, (b) 10% Ag/CeO2 (Imp), and (c) 10% Ag/CeO2 (DP). Reprinted from [41] with the permission of the Elsevier.
Catalysts 08 00285 g014
Figure 15. Temperature dependence of the formaldehyde conversion over the catalysts: ●, Ag/SiO2; ○, Ag/CeO2/SiO2; ◄, Ag/MnOx/SiO2; □, Ag/CeO2–MnOx/SiO2. Reaction conditions: 18,000–22,000 ppm of CH2O in dry air; catalyst mass 145 mg; WHSV = 69,000 mL/(gcath). Reproduced from [145] with permission from Elsevier.
Figure 15. Temperature dependence of the formaldehyde conversion over the catalysts: ●, Ag/SiO2; ○, Ag/CeO2/SiO2; ◄, Ag/MnOx/SiO2; □, Ag/CeO2–MnOx/SiO2. Reaction conditions: 18,000–22,000 ppm of CH2O in dry air; catalyst mass 145 mg; WHSV = 69,000 mL/(gcath). Reproduced from [145] with permission from Elsevier.
Catalysts 08 00285 g015
Figure 16. Structure of O2 adsorbed on Ag5/CeO2−x(111) surface complex. (a) Isomer where O2 is above the Ag cluster and forms a superoxo species (less stable); (b) isomer where O2 is below the Ag cluster and forms a peroxo group (more stable). Reproduced from Ref. [152]. Copyright© 2011, Elsevier.
Figure 16. Structure of O2 adsorbed on Ag5/CeO2−x(111) surface complex. (a) Isomer where O2 is above the Ag cluster and forms a superoxo species (less stable); (b) isomer where O2 is below the Ag cluster and forms a peroxo group (more stable). Reproduced from Ref. [152]. Copyright© 2011, Elsevier.
Catalysts 08 00285 g016
Figure 17. (a) Scanning electron microscopy image and (b) TEM image of the CeO2 nanopetaled rose-flower-like morphology annealed at 300 °C for 3 h. Insets present a high-resolution TEM image and a selected-area electron diffraction pattern of the CeO2 roses. Reproduced from Ref. [172] with the permission from the American Institute of Physics.
Figure 17. (a) Scanning electron microscopy image and (b) TEM image of the CeO2 nanopetaled rose-flower-like morphology annealed at 300 °C for 3 h. Insets present a high-resolution TEM image and a selected-area electron diffraction pattern of the CeO2 roses. Reproduced from Ref. [172] with the permission from the American Institute of Physics.
Catalysts 08 00285 g017
Figure 18. The proposed mechanism for the enhancement of photocatalytic activity of Ag2O/CeO2 catalyst in degradation of enrofloxacin. Reproduced from Ref. [167] with the permission from the Elsevier.
Figure 18. The proposed mechanism for the enhancement of photocatalytic activity of Ag2O/CeO2 catalyst in degradation of enrofloxacin. Reproduced from Ref. [167] with the permission from the Elsevier.
Catalysts 08 00285 g018
Figure 19. Schematic diagram showing: (a) transfer of electrons to form a Schottky barrier (b) transfer of electrons excited by surface plasmon resonance, (c) excitation of electrons in the photocatalyst from the local electric field. Reproduced from Ref. [170] with the permission from the Elsevier.
Figure 19. Schematic diagram showing: (a) transfer of electrons to form a Schottky barrier (b) transfer of electrons excited by surface plasmon resonance, (c) excitation of electrons in the photocatalyst from the local electric field. Reproduced from Ref. [170] with the permission from the Elsevier.
Catalysts 08 00285 g019
Table 1. A selection of CeO2-based catalysts for soot oxidation.
Table 1. A selection of CeO2-based catalysts for soot oxidation.
CatalystPreparation MethodCeO2 MorphologySBET, m2/gParticle Size, nmCe3+/Ce4+ RatioCatalyst/Soot RatioContact ModeReaction ConditionsT10, °CT50, °CT90/Tmax, °CRef.
AgCeO2
CeO2-NChydrothermalnanocubes11-1000.454:1
(mass.)
TC1% O2/N2 500 mL/min, isothermal reactions at 300 °C and 350 °C-430-[105]
CeO2-NPthermal decompositionirregular shaped71-150.57-458-
CeO2-Sphydrothermalspindles79-250.53-527-
CeO2-30precipitation at 30 °Cirregular shaped49-110.524:1
(mass.)
LC20% O2/80% N2--598[106]
CeO2-50precipitation at 50 °C41-150.51-542-
CeO2-70precipitation at 70 °C49-150.50-542-
Ce-Rhydrothermalnanorods80-250 nm × 2 μmCe3+ 25.1 at. %9:1
(mass.)
LC10 vol%O2/N2356500554[96]
TC286368400
Ce-Psolvothermalirregular shaped88-30–40Ce3+ 16.5 at. %LC413521573
TC320433474
Ce-Fsolvothermalflakes62-25Ce3+ 19.1 at. %LC433554622
TC306383440
Ce-SAShydrothermal route in a batch stirred-tank reactorSA stars124-10N/A45:5
(mass.)
LC50% air/ 50% N2 constant 100 mL min−1450560610[107]
TC385415505
Ce-NChydrothermalnanocubes4-54N/A45:5
(mass.)
LC 420465575[93]
TC370385430
Ce-NDthermal decompositionirregular shaped72-7–35N/A45:5
(mass.)
LC50% air/ 50% N2 100 mL min−1475530600
TC360390498
Ce-NChydrothermalnanocubes4-54Ce3+ 27.6 at. %45:5
(mass.)
LC10% of O2/N2 at rate of 100 cm3 min−1417477584[88]
TC396400425
Ce-NRhydrothermalnanorods4-43Ce3+ 25.5 at. %LC429536623
TC381416455
Ce-Mimproved graftingmesoporous75-5Ce3+ 25.5 at. %LC398538604
TC374464510
Ce-SCSsolution combustionmesoporous69-35Ce3+ 36.1 at. %LC436580633
TC392476558
CeO2-CP1-Fco-precipitationirregular shaped52.6-8.46Ce3+ 21.71 at. %45:5
(mass.)
LC5% O2/Ar, 200 mLmin−1--545[90]
CeO2-CP2-Fmodified co-precipitation HNO3/Ce(NO3)3 = 0.5 (mol)22.7-7.87Ce3+ 12.77 at. %--530
CeO2-CP3-Fmodified co-precipitation HNO3/Ce(NO3)3 = 1 (mol)24.6-6.05Ce3+ 11.90 at. %--480
CeO2-CP4-Fmodified co-precipitation HNO3/Ce(NO3)3 = 2 (mol)irregular shaped30.13-6.07Ce3+ 10.58 at. %45:5
(mass.)
LC5% O2/Ar, 200 mLmin−1--465[90]
CeO2-CP4-ACeO2-CP4 calcined at 750 °C for 6 hirregular shaped1.80-47.18Ce3+ 15.60 at. %--440
CeO2-S-Fsolid combustion77.1-9.63Ce3+ 26.58 at. %--540
CeO2-CA-Fcitric acid sol–gel45.0-9.68Ce3+ 30.76 at. %--560
CeO2-500electrospinning with calcination at 500 °Cnanofibers20.4-241–253N/A95:5
(mass.)
LC21% O2 and 79% N2, 100 mL/min 596[47]
TC--429
CeO2-800electrospinning calcination at 800 °C3.45-241–253N/ALC--633
TC--504
CeO2-1000electrospinning calcinations at 1000 °C3.40-241–253N/ALC--639
TC
--513
CeO2precipitationirregular-shaped45-N/AN/A20:1
(mass.)
TC10% O2/N2 10 °C min−1--393[48]
CeO2precipitation/ripeningnanofibers4-72N/A45:5
(mass.)
LC10% O2/N2480555560[92]
TC383439445
CeO2solution combustionuncontrolled nanopowders31-45N/ALC483562562
TC358411417
CeO2hydrothermalthree-dimensional SA stars105-9N/A45:5
(mass.)
LC10% O2/N2435543552[92]
TC354410403
CeO2SA stars aged 5 h at 600 °CAged SA stars50-15N/ALC473559559
TC381453465
AgCe-NCincipient wetness impregnation (Ag-5 wt. %)nanocubes101.5–3.51000.344:1
(mass.)
TC1% O2/N2 500 mL/min 100,000 h−1, isothermal reactions at 300 °C-376-[105]
AgCe-NPincipient wetness impregnation (Ag-5wt. %)irregular shaped641.5–3.5160.52-389-
AgCe-Spincipient wetness impregnation (Ag-5 wt. %)spindles691.5–3.5270.37-411-
Ag/CeO2-30incipient wetness impregnation (Ag-5 wt. %)irregular shaped375150.294:1
(mass.)
LC1% O2/N2 after 4 cycles522606691[106]
Ag/CeO2-50338200.27488596660
Ag/CeO2-70378150.23504602675
Ag/CeO2-500electrospinning calcination at 500 °C (Ag-4.5 wt. %)nanofibers5.0710241–253N/A95:5
(mass.)
LC21% O2, 79% N2, 100 mL/min--481[47]
TC--429
Ag/CeO2-800electrospinning calcination at 800 °C (Ag-4.5 wt. %)3.0710241–253N/ALC--485
TC--484
Ag/CeO2-1000electrospinning calcinations at 1000 °C (Ag-4.5 wt. %)2.7410241–253N/ALC--514
TC--496
Ag/CeO2incipient wetness impregnation (Ag-10wt. %)irregular shapedN/AN/AN/AN/A20:1
(mass.)
TC10% O2/N2 10 °C·min−1--345[48]
CeO2–Agco-precipitation (Ag-39 wt. %)rice-ball14.73616N/A19:1
(mass.)
LC10% O2/He at 50 mL/min--376[49]
TC315
Ag(39)/CeO2impregnation (Ag-39 wt. %)irregular shaped30.18921N/ALC--563
TC381
Ag(10)/CeO2impregnation (Ag-10 wt. %)52.06020N/ALC--526
TC362
Ag(3.2)/CeO2impregnation (Ag-3.2 wt. %)59.22820N/ALC--550
TC371
Ag(1.9)/CeO2impregnation (Ag-1.9 wt. %)70.02020N/ALC--596
TC414
Ag(0.95)/CeO2impregnation (Ag-0.95 wt. %)78.1n.d20N/ALC--610
TC466
Table 2. Literature data on catalytic VOCs abatement over Ag/CeO2 composite catalysts.
Table 2. Literature data on catalytic VOCs abatement over Ag/CeO2 composite catalysts.
Type of VOCPreparation MethodLoading of Ag, wt. %TVOC conv., °CS, m2/gMean Ag NP Diameter (nm)Reaction ConditionsTOF × 103, s−1T, °CRef.
CH2OCP61.3
28.4
15
7.69
80%: 15040.5 to 84.4N/A1 mL of catalyst, CH2O: 0.42%, methanol: 0.074%,
H2O: 19.9%, N2: 62.7%,
O2: 16.9% GHSV = 21,000 h−1
Trange: 423–573 K
--[140]
CH2OWI8100%: 125113.7N/A110 ppm of CH2O 20% O2, N2 balance
GHSV = 100,000 mL (gcat·h)−1 Kinetic studies: 1400 ppm of CH2O.
GHSV = 302,000 mL (gcat·h)−1
6.8100[137]
CH2OWI1100%: 10070.8<350 mg of catalyst 600 ppm CH2O
20.0 vol% O2, N2 balance GHSV = 120,000–360,000 h−1
1.8100[138]
CH2OHT
WI
2100%: 110HT: 125.4.
WI: 55.5
nanospheres
HT: 14.8
WI: 2.4
50 mg of catalyst powder mixed with quartz sand 810 ppm of CH2O 20% O2, N2 balance GHSV = 84,000 h−15.0110[40]
CH2OHT
WI
5100%: 110 (nr)
50%:
nr: 74
np: 89
nc: 108
nr: 128.46
np: 104.74
nc: 72.63
np: 4.0
nr: 6.0 ± 2.0 nm and 50.0–100.0 nm
50 mg of catalyst 810 ppm of CH2O
GHSV = 84,000 h−1 contact time was 0.34 s
Trange: 30–240
1.9 *
TOFAg
nr: 71.0
np: 46.0
nc: 31.0
100[139]
propyleneWI
DP
1050%:
WI: 173
DP: 261
WI:92
DP:84
N/A100 mg of fine catalyst powder, air and
6000 ppm of C3H6, reactive flow of 100 mL min−1
WI: 2.2
DP: 0.13
170[41]
propyleneWI
DPU
IRC
450%:
WI: 221
DPU: 260
IRC: 200
WI: 149
DPU: 123
IRC: 99
N/A200 mg of catalyst
6000 ppm of C3H6
a total flow of 100 mL/min
Trange: 60–400
WI: 0.27
DPU: 0.22
IRC: 0.14
170[141]
propyleneWI
DP
2.1450%:
WI: 220
DP: 245
WI: 98
DP: 118
N/A100 mg of fine catalyst powder
6000 ppm of C3H6
a total flow of 100 mL min−1
Trange: 100–400
WI: 0.8
DP: 0.5
170[142]
toluene 50%:
WI: 240
2000 ppm of C7H80.34170
tolueneDP
CP
4.8
4.7
50%:
DP: 265
CP: 260
DP: 112
CP: 130
DP: 7.1–6.7
CP: <4–3.3
0.7 vol.% VOC
10 vol.% O2
He balance
GHSV = 7.6 × 10−3 molVOC h−1 gcat −1
--[54]
methanol 50%:
DP: 131
CP: 113
--
acetone 50%:
DP: 225
CP: 220
--
naphthaleneWI1100%: 240
50%: 175 (1 wt. % Ag)
1438.7120 ppm naphthalene
10% O2, N2 balance
total gas flow rate was 400 mL/min
GHSV = 175,000 h−1
Trange: 160–300
1.5170[143]
WI—wetness impregnation, DP—deposition–precipitation, DPU—deposition–precipitation with urea, IRC—impregnation–reduction with citrate, HT—hydrothermal synthesis, nr—nanorod, np—nanoparticle, nc—nanocube. *—the calculation was carried out as a ratio of mole of converted formaldehyde per mole of Ag loading in the catalysts.

Share and Cite

MDPI and ACS Style

Grabchenko, M.V.; Mikheeva, N.N.; Mamontov, G.V.; Salaev, M.A.; Liotta, L.F.; Vodyankina, O.V. Ag/CeO2 Composites for Catalytic Abatement of CO, Soot and VOCs. Catalysts 2018, 8, 285. https://doi.org/10.3390/catal8070285

AMA Style

Grabchenko MV, Mikheeva NN, Mamontov GV, Salaev MA, Liotta LF, Vodyankina OV. Ag/CeO2 Composites for Catalytic Abatement of CO, Soot and VOCs. Catalysts. 2018; 8(7):285. https://doi.org/10.3390/catal8070285

Chicago/Turabian Style

Grabchenko, M. V., N. N. Mikheeva, G. V. Mamontov, M. A. Salaev, L. F. Liotta, and O. V. Vodyankina. 2018. "Ag/CeO2 Composites for Catalytic Abatement of CO, Soot and VOCs" Catalysts 8, no. 7: 285. https://doi.org/10.3390/catal8070285

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