Next Article in Journal
Preparation of Catalysts from Renewable and Waste Materials
Next Article in Special Issue
Development of Pharmaceutical VOCs Elimination by Catalytic Processes in China
Previous Article in Journal
The Spinning Voltage Influence on the Growth of ZnO-rGO Nanorods for Photocatalytic Degradation of Methyl Orange Dye
Previous Article in Special Issue
Comparative Study of ZnO Thin Films Doped with Transition Metals (Cu and Co) for Methylene Blue Photodegradation under Visible Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bimetallic Catalysts for Volatile Organic Compound Oxidation

Department of Chemical Sciences, University of Catania, Viale A. Doria 6, 95125 Catania, Italy
Submission received: 25 May 2020 / Revised: 7 June 2020 / Accepted: 9 June 2020 / Published: 12 June 2020

Abstract

:
In recent years, the impending necessity to improve the quality of outdoor and indoor air has produced a constant increase of investigations in the methodologies to remove and/or to decrease the emission of volatile organic compounds (VOCs). Among the various strategies for VOC elimination, catalytic oxidation and recently photocatalytic oxidation are regarded as some of the most promising technologies for VOC total oxidation from urban and industrial waste streams. This work is focused on bimetallic supported catalysts, investigating systematically the progress and developments in the design of these materials. In particular, we highlight their advantages compared to those of their monometallic counterparts in terms of catalytic performance and physicochemical properties (catalytic stability and reusability). The formation of a synergistic effect between the two metals is the key feature of these particular catalysts. This review examines the state-of-the-art of a peculiar sector (the bimetallic systems) belonging to a wide area (i.e., the several catalysts used for VOC removal) with the aim to contribute to further increase the knowledge of the catalytic materials for VOC removal, stressing the promising potential applications of the bimetallic catalysts in the air purification.

Graphical Abstract

1. Introduction

Volatile organic compounds (VOCs) are a wide group of organic compounds characterized to boiling points less than 250 °C at room temperature and at atmospheric pressure [1]. Due to their carcinogenic and toxic nature, most VOCs are considered major causes of air pollution. Indeed, their emission in the environment leads to the formation of secondary dangerous compounds, due to the occurrence of chemical reactions with other airborne pollutants such as NOx and SOx, which results in the formation of tropospheric ozone and photochemical smog [2,3]. Long exposure to these pollutants leads to serious problems for human health [4,5]. Global economic and industrial development over the years has caused an exponential increase of anthropogenic VOC emission [6]. VOC discharges include outdoor sources, such as transport, industrial and petrochemical processes, etc., and indoor sources, such as household products, solvents, office materials, cleaning products, domestic cooking, etc. [7]. The emitted VOCs encompass alkanes, paraffins, olefins, aromatics, alcohols, ketones, aldehydes, esters, sulfur/nitrogen-containing VOCs, and halogenated VOCs. Among them, the most common and toxic are benzene, phenol, toluene, styrene, formaldehyde, propylene, and acetone [8], whereas Cl-VOCs and in general halogenated VOCs, due to their inherent stability and toxicity, are also very dangerous [9].
Different technologies have been developed for VOC treatment, and they can be divided into nondestructive and destructive VOC removal. The former include adsorption, membrane separation, and condensation [10,11,12]. Among these, the adsorption process is considered one of the most efficient treatments, owing to a low energy consumption, relatively low operation cost, and simple operations for the adsorption/regeneration of the adsorbent [12,13]. With adsorption, it is possible to remove, without the generation of dangerous byproducts, a low/medium concentration of VOCs (<1000 mg/m3) [1]. Regarding destructive (i.e., oxidative) processes, the most commonly used ones are thermal (not-catalytic) combustion and catalytic oxidation, both of which can be applied to treat a medium/high concentration (>5000 mg/m3) of VOCs [14,15]. In particular, catalytic conversion has some advantages compared to thermal incineration; indeed, this process has become more popular than noncatalytic treatments. Catalytic oxidation allows converting VOCs into less toxic substances, such as carbon dioxide and water, in a temperature range much lower than thermo-oxidation [16,17]. Specifically, with catalytic conversion, the operating temperature range is 200–500 °C or even lower, whereas in thermal incineration, temperatures are higher (800–1200 °C). Lower temperatures permit reducing the production of dioxins and NOx. Furthermore, catalytic oxidation is more versatile and cheaper, especially when it comes to processing low concentrations of organic compounds [18]. In recent years, new technologies have been applied for the elimination of VOCs at low concentrations, namely, the advanced oxidation process (AOP), e.g., photocatalytic degradation, ozone treatment, Fenton oxidation methods [19,20], biodegradation [21], and phytoremediation [22].
Due to the economic and technological advantages of catalytic oxidation, widespread efforts have been committed to the selection of high-performing catalysts for this process. However, considering the large number of organic molecules and the problematic nature of VOCs mixtures, the design and optimization of catalytic materials are challenging tasks. Both noble and transition metals have been widely used as catalysts for either nonhalogenated or halogenated VOCs [23,24,25]. Notwithstanding their high costs, the supported noble-metal catalysts are widely applied due to their intrinsic features, such as resistance to deactivation, ease of regeneration, and highly catalytic performance [26,27,28]. These features strictly depend on the synthetic procedure adopted for the preparation of the supported metal catalyst, as well as the type of metal salt precursor, the metal loading, the kind of support, and the particle size [29,30,31]. Furthermore, VOCs and air/oxygen content, total gas stream rate, and employed reactor (membrane reactor, fixed-bed reactor, etc.) are key parameters that can affect overall catalytic activity [32,33,34].
In the literature, there are many studies that deeply analyze single or various parameters that influence the final results of catalytic oxidation applied to VOC treatment, including the catalysts used [16,35,36], the nature of VOCs [2,9,37], the combination of different technologies [1,38], the type of reactor [39], or the performances in practical applications [40].
One of the less explored strategies to enhance the catalytic activity of supported noble/transition metal catalysts is the addition of a second metal (noble and/or transition) to the first one.
This work analyzes a little aspect of the VOC catalytic treatment topic: the advantages of using supported bimetallic catalysts with respect to monometallic counterparts, focusing on the morphological, chemicophysical, and textural properties of these peculiar materials, and how these features can influence the catalytic activity.
This review aims to enlarge the scientific panorama about VOC removal through catalytic oxidation, focusing on the bimetallic catalysts, an aspect not yet systematically examined in the literature.

2. Bimetallic Catalysts for VOC Oxidation

Bimetallic nanoparticles (NPs) are a kind of materials formed by two different metals and characterized with peculiar features [41,42]. Specifically, they can show new properties resulting from the combination of features arising from the monometallic counterparts. Usually, the obtained physicochemical properties of bimetallic systems give a holistic result, i.e., the final properties are not the simple additive features of the monometallic analogs, but in many cases, it is possible to exploit a great improvement with new properties due to the presence of synergistic effects [43,44]. Since their application in the field of petrochemistry [45], bimetallic systems are being widely applied in heterogeneous catalysis in various reactions, such as hydrogenation [46,47], reforming [48,49], H2 production and purification [50,51,52,53], and oxidation [54,55,56]. Recently, they have also been applied in the biomedical field [57].
On the basis of the morphology of the bimetallic system, it is possible to classify the main arrangements into three typologies: core–shell or multishell structures, heterostructures, and random or homogeneous alloys (Figure 1) [43,44,58].
Various factors influence the final morphology of bimetallic systems: (1) the intrinsic strength of the bonds between the two metals in comparisons to the strength of the bonds between the two monometallic constituents (in particular, if the resulting alloy bond strength is greater, reciprocal mixing is preferred, and, in the opposite case, the segregation of the two metals is favored); (2) the balance in the surface energy of the metals, where the monometallic element that owns the lower surface energy will move to the surface, establishing a shell structure, while the metal with the smaller atomic sizes will collocate to the core; (3) the mixing is favored when an electron/charge transfer between the metallic counterparts is verified; and (4) magnetic effects can influence the final structure between the two metals [44,58,59].
The morphologies are also strictly affected by the preparation method adopted. In general, the synthesis of the bimetallic systems can be carried out with the solid-state or with the solution methods [60,61,62]. The latter are preferred because solid-state techniques require high temperatures and long times for annealing procedures, thus decreasing the surface area of the bimetallic catalyst and affecting catalytic performance in a crucial way. By contrast, with solution methods, it is possible to control the nucleation and growth processes by modifying the reaction parameters.
In the main catalytic reactions, the bimetallic system is supported on a specific support [63]. The commonly used techniques for the preparation of supported bimetallic catalysts are: impregnation [64,65], co-precipitation [66,67], deposition–precipitation [50,68], thermal decomposition [69], and chemical reduction [70]. The choice of a proper support is another key feature that strongly influences the catalytic activity of the bimetallic system. In particular, for VOC oxidation, and for oxidation reactions in general, the utilization of an active support (i.e., having redox properties and lattice oxygen mobility), for example, CeO2 or Fe2O3, can contribute to the formation of a synergistic effect involving both the metals and the same support [68,71], whereas other supports, with high surface areas, such as zeolites, silica, alumina, etc., favor the high dispersion of the metal species and, therefore, the interaction between the metals and the same support [16,72].
Since the first works of Haruta and co-workers [73], gold-based catalysts have generated many investigations focused on the catalytic properties of this peculiar metal. With respect to the platinum-group catalysts, gold’s properties are in some cases superior; consequently, the supported gold-based materials have found many applications in catalytic reactions dealing with environmental protection and energy production [50,72]. In the specific field of VOC catalytic oxidation, gold-based catalysts play a key role. For these reasons, bimetallic alloys with gold are firstly examined.

2.1. Gold-Based Bimetallic Catalysts

The important role of gold-based catalysts in heterogeneous catalysis was disclosed through the high activity of this metal in CO oxidation and selective oxidations [74,75]. After these studies, many works dealing with this catalyst have emerged, widely expanding the use of gold catalysts [72,76]. With respect to the other commonly used noble metals, namely, Pt and Pd, gold catalysts show peculiar features such as resistance to O2 poisoning and high selectivity. Another key parameter that strictly affects the catalytic activity of Au catalysts is size-dependence: To be active, gold particles must be, in general, smaller than 5 nm, which balances the proportion of the low coordination surface active sites (edges and corners) [77].
For other applications, as well as for VOC oxidation, the catalytic activity of gold-supported samples is affected by many factors: (a) gold–support interaction; (b) gold loading; (c) the valence state of gold; (d) the adopted synthesis and pre-treatment conditions utilized; and (e) the concentration and nature of the chosen VOC target [78,79]. Among the monometallic gold catalysts employed for VOC oxidation, the Au/CeO2 sample showed a great performance in the oxidation of oxygenated molecules, i.e., aldehydes, ketones, esters, and alcohols [72]. It was reported that the high activity was related to the enhancement in the surface oxygen reducibility/mobility of the CeO2 support. The oxidation mechanism followed a Mars–Van Krevelen (MvK) mechanism, where the ceria lattice oxygens were actively involved in the oxidation pathway [68,80] (Figure 2).
Nevertheless, the application of gold catalysts to industrial level is limited due to certain drawbacks, such as the decrease of activity in the presence of high concentration of moisture and the aggregation of gold nanoparticles at high temperatures [81,82].
Gold-based bimetallic systems are considered a useful solution to overcome the cited limitations, due to the combination of properties from the gold and the second metal. Among the various metals employed to be joined with gold, it has been found that Pd, Ag, and Cu show a high miscibility, and these systems can be prepared with a wide range of methodologies [83,84], whereas Pt, Ru, Co, Fe, and Ni are only partially or not at all miscible with gold [84]. Consequently, Au-Pd followed by Au-Ag and Au-Cu are the most utilized gold-based bimetallic catalysts for VOC oxidations.

2.1.1. Au-Pd Catalysts

Gold is miscible with palladium in all compositions; consequently, while the formation of gold–palladium alloys is favored, the segregation of single metals was, in fact, avoided [76,83]. In Table 1 are reported some of the experimental results of the application of the Au-Pd systems in catalytic oxidation of various VOCs (T90 = temperature at which the 90% of conversion was achieved).
In particular, Hosseini et al. [85] synthetized Au-Pd catalysts supported on mesoporous TiO2 for the removal of toluene, propene, and a gaseous mixture of both. Interestingly, they found that catalytic activity is influenced by the morphology of the core–shell structure with the best performance shown with the Au-core/Pd-shell. By contrast, with the reverse morphology (Au-shell/Pd-core) the catalytic activity was lower, due to the lower affinity of gold for oxygen adsorption (in this case, the rate determining step of the reaction that followed a Langmuir–Hinshelwood mechanism) caused by the poor ability of gold to polarize oxygen molecules. In the same context, Barakat et al. [86] investigated the catalytic stability of bimetallic Au-Pd/doped TiO2 samples under severe testing conditions (exposing the catalyst to 110 h of a gaseous toluene/air stream). The bimetallic catalyst maintained a good activity even after a long time. The interaction between the Nb-doped TiO2 support and the Au-Pd system allowed obtaining a cycle-like activity of the catalyst. This oscillatory behavior was related to the existence of carbonaceous compounds adsorbed on the surface of the spent catalyst that, together with the formed OH radicals, favored the reduction of palladium. The redox process of palladium was linked to the cyclic-like activity of the bimetallic sample (Figure 3).
Regarding toluene degradation, Xie et al. [88] studied the catalytic performance of the bimetallic gold–palladium system supported on three-dimensionally ordered macroporous (3DOM) Co3O4. The bimetallic sample showed a much higher activity compared to its nonmetallic counterparts, with a T90 of about 100 and 30 °C lower compared to monometallic gold and palladium, respectively. The peculiar features of the 3DOM supports (such as higher porosity and ordered pore channels) were also exploit by the same authors [91], who utilized an Au-Pd bimetallic sample prepared via chemical reduction, but employing the Mn2O3 as support. The bimetallic catalysts confirmed the higher activity compared to monometallic gold and palladium in the oxidation of different VOCs, such as methane and o-xylene. To further boost catalytic activity, doping with Fe of the Au-Pd/3DOM-Mn2O3 catalyst allowed modifying the structural properties of the alloy NPs. With this modification, oxygen activation and the methane adsorption ability were increased, enhancing, as final result, the overall catalytic activity.
Catalytic performance in various mixtures of VOCs (toluene/m-xylene, ethyl acetate/m-xylene acetone/m-xylene, and acetone/ethyl acetate) was examined by Xia et al. [87] on Au-Pd prepared via chemical-reduction-supported αMnO2 nanotubes. In this case, the authors focused on an MvK-like mechanism with the mutual interaction of the Au-Pd nanoparticles with the α-MnO2 that improved the mobility/reactivity of the surface lattice oxygen of the support. In catalytic oxidation of VOC mixture, the rate-determining step is the competitive adsorption between the various VOCs on the surface of the catalyst. With the bimetallic catalyst, the authors measured the total oxidation of a single component and of the VOC mixture at T < 300 °C. The same bimetallic sample also showed a high catalytic stability in the long time (50 h) on stream experiments.
Tabakova et al. [89] focused their work on the removal of benzene utilizing an Au-Pd bimetallic system synthetized by deposition–precipitation and supported on Fe-doped CeO2. The superior performance of the bimetallic catalyst with respect to the monometallic ones was highlighted through comparison of the T90, which was ≈95 °C for the bimetallic sample, ≈180 °C for the Pd/Fe-CeO2, and ≈190 °C for the Au/Fe-CeO2. In addition, in this case, the synergistic interaction between the alloy nanoparticles and the support enhanced the mobility of the ceria surface lattice oxygen, further boosted by doping with iron. Moreover, this strong interaction facilitated the nucleation of the noble metal particles on the surface of cerium oxide.
Benzyl alcohol oxidation on Au-Pd bimetallic catalysts was extensively studied by the research group of Prati et al., and an exhaustive comparison of the performance, in this reaction, of the bimetallic Au-Pd systems present in the literature, together with a deep analysis of the various adopted preparation methods are reported in [83]. The same research group [76,83] studied the catalytic behavior of different bimetallic systems composed of Au and various other metals supported on activated carbon prepared using the sol immobilization methodology. Interestingly, they found a structural correlation depending on the second metal utilized. Specifically, an alloy structure was obtained using Pd and Pt, whereas a core–shell morphology was attained with Ru, while with Cu, a phase segregation of this metal, instead of gold, was favored. In the benzyl alcohol tests, the bimetallic synergistic effect was exploited only with copper and palladium.
The strong interaction between gold and palladium, beneficial for catalytic oxidation of benzyl alcohol, was also examined by Li et al. [90] using CeO2 with different morphology as support. The different morphology of cerium oxide (rod, cube, and polyhedrons), where gold/palladium nanoparticles were deposited–precipitated (Figure 4), affected the catalytic performance. Specifically, the Au-Pd supported on ceria rod showed a higher benzyl alcohol conversion with respect to the samples supported on CeO2 cubes and CeO2 polyhedrons and was thus related to the smaller particle size of ceria rod compared to the other CeO2 supports; moreover, this particular morphology also favored a higher concentration of ceria oxygen defects, enhancing the mobility/reducibility of the ceria surface oxygens. By contrast, the sample supported on the CeO2 cube exhibited the highest selectivity in benzaldehyde.
Kucherov et al. [92] investigated the performance of mono- and bimetallic gold-based catalysts for the removal of dimethyldisulfide (DMDS), an S-VOC. They corroborate the performance of Au-Pd catalysts also for this type of VOC. The bimetallic sample supported on TiO2 demonstrated a stable performance and assisted with the removal of DMDS at T < 155 °C with the formation of SO2 and elemental S.

2.1.2. Au-Ag and Au-Cu Catalysts

The establishment of a strong interaction between gold and silver with the formation of an alloy or of bimetallic clusters was investigated by our research group both in VOC oxidation and in H2 purification towards the preferential oxidation of CO (PROX reaction) [50,53,68]. In particular, we evaluated the catalytic activity of Au-Ag and Au-Cu bimetallic samples supported on CeO2 toward the degradation of 2-propanol and ethanol. A higher activity was found of the gold–silver sample with respect to Au-Cu and the monometallic counterparts. The higher activity of the gold–silver system was correlated to a higher mobility/reactivity of ceria surface oxygens, due to a strong synergistic interaction between the gold–silver nanoparticles and the cerium oxide. A linear correlation was stated considering the T50 of alcohol oxidation and the TPR (temperature-programmed reduction) initial temperature, i.e., the temperature at which the reduction of ceria surface oxygens started considering the analyzed samples (Figure 5). The Au-Ag/CeO2 catalyst displayed the lowest reduction temperature and T50.
Nagy et al. [93] studied the performance of Au-Ag nanoparticles supported on SiO2 synthetized from the adsorption of bimetallic colloids in the oxidation of benzyl alcohol. The authors focused their research on the crucial importance of the molar ratio between the two metals. In particular, a synergistic effect was verified that reflects a higher activity at a low Ag/Au molar ratio (best result Ag/Au = 23/77). For the authors, the synergy is activated by the optimal concentration of the two metals, which increased the activation of both oxygens from gas-phase and from the support. In the same context, a correlation between catalytic activity and the concentration of gold and silver was measured by our research group in the PROX reaction [50] with a higher concentration of gold or silver with respect to the second metal that was detrimental for the overall catalytic performance, whereas the best results were obtained with an approximately equal concentration of gold and silver (1% wt–1% wt).
The crucial importance of the molar ratio between gold and the other metal was also stated in the review of Bracey et al. [94], focused on the Au-Cu system. Specifically, in one of the analyzed works, the following order of reaction in catalytic oxidation of propene is reported: AuCu (1:3 molar ratio)/TiO2 > AuCu (1:1 molar ratio)/TiO2 > AuCu/TiO2 (3:1 molar ratio) > Au/TiO2. The content of copper, in fact, strongly influenced the dispersion of the metal nanoparticles, with a high amount of copper in the alloy that caused a decrease in the size of the metal particles, thus contributing to enhancint the activity and selectivity into propene oxide [95]. In the same review, it was illustrated that, when investigating another reaction, such as selective oxidation of benzyl alcohol to benzaldehyde, the more active bimetallic catalyst was the sample with the higher concentration of gold (the catalyst AuCu/SiO2 with a molar ratio of 4:1). Similarly to the previous examples, the bimetallic catalyst was prepared by impregnation, but in this case, a higher concentration of gold is fundamental to achieve a high selectivity (98%) to benzaldehyde.
The above-discussed literature data on AuCu bimetallic catalysts were mainly focused on the selective oxidation of VOCs, whereas the work of Nevanperä et al. [96] dealt with catalytic oxidation of DMDS with bimetallic gold-based catalysts (Au-Cu and Au-Pt) supported on γ-Al2O3, CeO2, and CeO2-Al2O3 prepared by surface redox reduction. Among the examined supports, the alumina gave the best results, whereas the addition of gold enhanced the catalytic activity of both monometallic copper and platinum samples, Au-Cu catalysts being the most active system. Interestingly, the authors noted that the same Au-Cu catalyst led to the formation of dangerous byproducts, such as carbon monoxide and formaldehyde. This was attributed to the high concentration of reactive surface oxygens favored by the presence of copper oxide and to the dissociation of the oxygen that started at a lower temperature with respect to the monometallic samples, with the consequent modification of the surface acid and basic sites of the bimetallic catalyst. By contrast, selectivity towards CO2 and H2O was higher in the Au-Pt sample.

2.1.3. Other Au-Based Bimetallic Catalysts

As discussed in the last examined work, among the other Au-based bimetallic catalysts, the Au-Pt system exhibited promising performance in VOC oxidation [96,97,98].
Kim et al. [97] investigated catalytic oxidation of toluene employing the Au-Pt/ZnO-Al2O3 catalyst prepared by impregnation in air or H2. They found that the bimetallic sample prepared in air led to an increase of the gold particle size and a decrease of the Pt with respect to the same particles synthetized in H2 stream, where an inverse correlation was verified (the gold size decreased, and the platinum size increased). Due to the crucial importance of the gold nanoparticles that facilitated the total oxidation of toluene and that increased the reduction of the surface oxygen of the mixed oxide support, the catalytic performance was higher with the bimetallic sample synthetized in H2 stream and calcined at 400 °C (Pt and Au mean size of about 5 nm). In another study [98], the same authors correlated the catalytic activity of the same bimetallic samples, even in the total oxidation of toluene, to the molar ratio of gold and platinum, finding the following order of activity: Pt75Au25 > Pt67Au33 > Pt100Au0 > Pt50Au50 > Pt33Au67 > Pt25Au75 > Pt0Au100. The small amount of gold promotes the total oxidation of toluene due to the formation of a strong metal–metal interaction.
The good affinity of gold with noble metals was also confirmed via the catalytic performance in VOC removal of the Au-Ru system [99,100]. Sreethawong et al. [99] investigated catalytic oxidation of methanol over gold–ruthenium samples prepared through impregnation and supported on SiO2. The characterization measurements (TPR, SEM, and XRD) suggested the occurrence of an interaction between the two metals exploited with a particular composition (3.32 wt%Ru–0.61 wt% Au), which led to obtaining a good catalytic activity, notwithstanding the fact that the two metals were not miscible in their bulk phase. Interestingly, if alumina was used as support, the formation of byproducts (methyl formate, formic acid, dimethyl ether, and formaldehyde) other to CO2 was detected, with an increase of methanol conversion. Catalytic oxidation of methanol was also analyzed by Calzada et al. [100] with Au-Ru/TiO2 catalysts prepared by deposition–precipitation with urea. The authors highlighted that the synergistic effect between the two metals was activated at a low conversion temperature (from room temperature to T = 50 °C), with a dependence on the bimetallic atomic ratio (the best-performing one was Ru:Au 0.75:1). Interestingly, the DRIFT (Diffuse Reflectance Infrared Fourier Transform Spectroscopy) CO spectra (Figure 6) illustrated as the interaction between the two metals decreased CO adsorption in the Ru surface sites in the Ru-Au 0.75:1 sample, an indication of the modification of the surface gold sites.
The modifications of the bimetallic surface sites due to interaction between the two metals strongly occurring with the Ru:Au 0.75:1 ratio were the reason behind the higher catalytic activity of this bimetallic sample compared to those of the monometallic catalysts and the Ru:Au 1:1 sample. In addition, in this case, the formation of formates as intermediates of the oxidation reaction was verified.
Catalytic activity in the total oxidation of toluene of the Au-Ir bimetallic catalyst supported on TiO2 was studied by Torrente-Murciano et al. [101]. Similarly to the previous case, the synergistic interaction between the two metals allowed them to sensibly decrease the T90 that was ≈230 °C for the bimetallic sample, ≈250 °C for the monometallic gold, and ≈270 °C for Ir/TiO2. The key factors that deeply influenced catalytic activity were the strong metals–support interaction exploited with the bimetallic system, which also permitted diminishing the loss of activity due to the metals sintering at high temperatures. Furthermore, the intimate contact between iridium and gold modified the bimetallic surface-active sites enhancing oxygen activation.
In the work cited in Section 2.2, Kucherov et al. [92] demonstrated a good activity of Au-Rh supported on HZSM-5 zeolite into the oxidation of the DMDS in SO2 at 290 °C. In this case, the zeolite support, owing to a high surface area, favored a high dispersion of the metals, a feature that is beneficial for catalytic activity.
Regarding gold–copper bimetallic systems, there are certain studies with other transition metals, utilized together with gold for VOC oxidation. Au-Co and Au-Fe interaction were principally investigated. In the examined works, the gold atoms interacted with the second metal present as a doping agent of the support [102,103,104].
Solsona et al. [102] synthetized gold nanoparticles anchored on cobalt containing mesoporous silica (UVM-7). The interaction between gold and cobalt permitted increasing catalytic activity in the oxidation of toluene and propane with respect to Au/UVM-7 and Co/UVM-7 catalysts. The presence of gold enhanced the reducibility of cobalt, present as Co3O4 at the Au-Co interface, thus facilitating the redox cycle of cobalt, with an MvK-like mechanism, which boosted catalytic oxidation of VOCs. In the same context, Albonetti et al. [103,104] deeply investigated the catalytic behavior of gold catalysts supported on mesoporous silica (SBA-15) via an iron oxide layer obtaining the Au/FeOx/SBA-15 composite. The good dispersion of nanosized gold favored the incidence of a strong synergism between gold and iron that led to an optimal activity in the combustion of methanol (T90 ≈ 140 °C).
As a conclusion of this generic overview of gold-based bimetallic catalysts applied at VOC oxidation, it is possible to recognize some fundamental features of these peculiar catalysts: (a) the essential action of the nanosized gold that is able to establish a metal–metal surface interaction with a wide range of both noble and transition metals; (b) the occurrence of a synergism between the two metals that allows sensibly decreasing the light-off temperatures of VOC oxidation; (c) the synergistic effect, which is not simply the addition of the single characteristics of the corresponding monometallic samples but leads to exploring new physicochemical properties; (d) the mutual interaction between the two metals which also strongly influences metals–support interaction (in particular, if the support is a reducible oxide (CeO2, MnOx, CoOx, etc.), the gold-based bimetallic cluster increases the mobility of the surface oxygen of the support, enhancing, in this way, catalytic oxidation towards an MvK mechanism). If the support is a nonreducible or hardly reducible oxide (TiO2, SiO2, and zeolites), the high dispersion of the bimetallic alloy and the modifications of the metals surface active sites allow enhancing oxygen adsorption, improving, as a final result, the overall catalytic performance. The most employed preparation methods of the supported gold-based bimetallic catalysts are impregnation (wet or wetness), deposition–precipitation, and chemical reduction.

2.2. Other Bimetallic Catalysts

Among the other noble metals, the most employed catalysts for catalytic oxidation of VOCs are platinum-based materials [16], and similarly to gold, platinum has shown a good affinity with palladium [105,106,107,108]. In general, as can be seen from Table 2, the use of noble-metals-based bimetallic catalysts has allowed obtaining a good performance in the removal of VOCs, whereas the utilization of transition-metals-based materials has led to shift at high temperatures in the total conversion of VOCs. However, especially in recent years, the necessity to reduce the amount of expensive noble metals has led to exploring a new synergism between noble and transition metals.
Fu et al. [105] prepared, for the hydrothermal method, a Pt-Pd bimetallic sample supported on mesoporous silica. Comparably to the catalytic behavior of gold-based samples, the synergism between the two metals allowed obtaining a superior performance in the removal of toluene with respect to the monometallic samples, with an improvement in the reducibility of palladium, involved in the redox cycle PdOPd0, and in oxygen adsorption capability.
The good catalytic activity of the Pt-Pd system was also confirmed by Kim et al. [106] in the degradation of benzene. Catalysts were synthetized via wetness impregnation on γ-Al2O3. Metal–metal interaction was favored by the formation of small and uniform particles and, as stated in the previous paragraphs, a specific amount ratio between the two metals (the optimum in the cited work is 0.3 wt% Pt–2%wt Pd). A higher concentration of platinum led to a remarkable decrease in activity due to the blockage of the active sites. The same authors in another paper [107] confirmed with a deep XPS (X-ray photoelectron spectroscopy) analysis the crucial role of the ratio between the metals to avoid the obstruction of the catalyst surface sites. The removal of methanol, acetone, and methylene chloride was instead studied by Sharma et al. [108], utilizing ceramic Raschig rings coated with Pt and Pd on fluorinated carbon. The authors measured a higher activity of the bimetallic catalyst with respect to monometallic ones. Furthermore, the hydrophobic nature of this particular bimetallic catalyst allowed obtaining a 90% of degradation of methanol and acetone at about 150 and 300 °C respectively, whereas 60% of degradation was achieved at 400 °C for the methylene chloride. In this case, a good correlation was established with a semi-empirical Langmuir–Hinshelwood model, which is able to predict the oxidation rate of each VOC in a gas mixture (methanol, acetone, and methylene chloride).
Ethanol adsorption and oxidation were investigated by the research group of Wittayakun et al. [114] with a Pt-M (M = Co, Cu, Mn) sample supported on silica MCM-41. Among the transition metals, cobalt gave the best results, and in particular, the bimetallic 0.5 wt% Pt–15 wt% Co exhibited the best ethanol adsorption and CO2 desorption. Interestingly, the authors identified two different reaction mechanisms considering the platinum monometallic sample and the bimetallic platinum–cobalt one (Figure 7).
Specifically, in the monometallic sample after the adsorption of ethanol, a formation was verified of a parallel adsorbed acetaldehyde, further converted into monodentate acetate and at end, dissociated and desorbed as carbon dioxide, methane, and water (Figure 7a). In the bimetallic catalyst, by contrast, the ethoxy species reacted with the adsorbed oxygen to give a bidentate acetate species that was transformed into carbon dioxide (Figure 7b). The modification of the ethanol adsorption led to a higher ethanol conversion with the monometallic platinum sample in comparison with the bimetallic Pt-Co sample that, conversely, showed a higher catalytic stability. Even with the bimetallic platinum-based catalysts, Chantaravitoon et al. [115] examined the performance of a Pt-Sn/γ-Al2O3 catalyst prepared with impregnation, for the oxidation of methanol. The authors noted from the temperature-programmed desorption (TPD) measurements of methanol oxidation that on the bimetallic catalyst, methanol decomposed as H2 and CO and the desorption peaks shifted at higher temperatures, increasing the amount of Sn. In addition, in this case, the monometallic Pt catalyst exhibited a better performance compared to the bimetallic one; however, the addition of a small amount of Sn (<5 wt%) reduced the deactivation of the catalyst in the long-time tests.
In addition, Ru-based bimetallic compounds were discreetly studied for VOC oxidation [109,110,116]. Liu et al. [109] prepared Ru-M (M = Co, Mn, Ce, Fe, Cu) samples supported on TiO2, evaluating catalytic performance in the degradation of benzene. Among the various metals, 1% wt Ru–5% wt Co showed the best activity; the presence of ruthenium, in fact, increased the reducibility of Co3O4. The authors stated also that the presence of water vapor inhibited benzene oxidation at T = 210 °C.
The total oxidation of propene was examined on Ru-Re/γ-Al2O3 by Baranowska et al. [116]. As discussed before, this nonconventional combination between these two metals also allowed increasing catalytic stability in the consecutive tests instead of the overall catalytic activity, which remained superior with the monometallic Ru sample. The addition of Re (the best composition being 5% wt Ru–3% wt Re) hampered the formation of RuO2 agglomerates. In this way, the dispersion of ruthenium is favored, allowing a higher stability compared to monometallic ruthenium catalysts. Ye et al. [110] performed a catalytic test regarding chlorobenzene removal with Ru-Ce/TiO2 samples prepared via impregnation. Interestingly, on the basis of the crystalline phase of titanium dioxide, the catalytic activity changed. At 280 °C, the bimetallic sample showed a conversion of 91% and 86% if supported on TiO2 rutile and TiO2 P25 (80% anatase, 20% rutile), respectively. The mixed crystalline phase of P25 was the best support for the monometallic ruthenium catalyst. By contrast, with respect to the work of Baranowska et al. [116] for this reaction and with the titanium dioxide support, dispersion was not the major parameter that affected catalytic activity; indeed, on 1%wt Ru–5%wt Ce/TiO2 (rutile), the abundant RuO2 clusters favored both catalytic activity and stability.
Interaction with Ce/CeO2 was also investigated by Yue et al. [117] but utilizing palladium. The performance of the bimetallic catalyst Pd-Ce/ZMS-5 synthetized through impregnation was evaluated on the degradation of methyl ethyl ketone (MEK). The presence of cerium oxide considerably increased the acid sites of palladium, enhancing at the same time the re-oxidation of Pd and boosting, in the end, the overall MEK degradation rate through an MvK mechanism.
Another bimetallic catalyst with palladium was prepared by Arias et al. [111]. In this work, the synergistic effect between palladium and manganese was explored utilizing alumina as support. The authors followed the oxidation of a VOC mixture (formaldehyde/methanol), concluding also in this case that an MvK-like mechanism was the reaction pathway, with the interaction between palladium and manganese favoring the oxidation of VOCs due to the activation of the reactive lattice oxygen of PdO and MnOx.
Various bimetallic samples were tested for catalytic oxidation of VOCs, studying both the physicochemical properties and the catalytic activity of silver-containing samples [113,118,119,120]. In particular, Jodaei et al. [113,118] tested different Ag-M bimetallic samples supported on ZMS-5 zeolite obtained via ionic exchange. The authors investigated the catalytic combustion of ethyl acetate, finding this order of activity and stability: Fe-Ag/ZSM-5 > Co-Ag/ZSM-5 > Mn-Ag/ZSM-5 > Ag/ZSM-5. The high dispersion of silver was favored by an optimal amount of iron (1.3 wt% Fe–1.75 wt% Ag), thus activating a synergistic effect between the two metals. In the same context, Izadkhah et al. [119] made a theoretical model for the removal of ethyl acetate. In particular, considering the preparation condition, the formulation, and loading of the promoter of silver, with their algorithm, it was possible to identify the optimal catalyst for this reaction. Among the first transition metal series, the bimetallic catalyst that exhibited superior performance compared to the monometallic silver was Fe-Ag/ZSM-5, thus confirming the experimental results of Jodaei et al. [113,118], Ni-Ag/ZSM-5 and V-Ag/ZSM-5.
Complete oxidation of formaldehyde at T < 90 °C was obtained by Qu et al. [120] with Ag-Co/MCM-41 silica. The key feature able to sensibly increase catalytic performance with respect to the monometallic silver was electron transfer between silver and cobalt that enhanced the reducibility of cobalt oxide, increasing, at the same time, the activation of surface oxygen on the bimetallic catalyst. Furthermore, the high metal–metal support interaction (SMMI) at the optimal Ag/Co mass ratio (3:1) favored a faster adsorption–dissociation of formaldehyde on the Ag species with respect to the Co3+ sites (Figure 8), thus decreasing the light-off temperature of VOC oxidation.
On the same support (MCM-41 silica), Pârvulescu et al. [121] synthetized with the hydrothermal method various Co-based bimetallic mesostructures (Co-V, Co-La, Co-Nb) characterized vy a high surface area and narrow pore size distribution. The oxidation of styrene and benzene was deeply influenced by the addition of a second metal component. Indeed, although the addition of La did not result in any synergistic effect, the addition of vanadium favored the oxidation of benzene, whereas the addition of niobium facilitated the removal of styrene, demonstrating that the presence of the second metal changed the surface-active sites of cobalt.
Similarly to cobalt, copper-based bimetallic catalysts also showed a good activity in the removal of VOCs [112,122,123]. Kim et al. [112] found an optimal interaction between Mn and Cu for the total oxidation of toluene. The order of activity considering other transition metals as a second component was: 5% wt Cu–15% wt Mn/γ-Al2O3 > 5% wt Co–15% wt Mn/γ-Al2O3 > 5% wt Ni–15% wt Mn/γ-Al2O3 > 15% wt Mn/γ-Al2O3 > 5% wt Fe–15% wt Mn/γ-Al2O3. The interaction between Mn and Cu favored a high dispersion of manganese, increasing, at the same time, the mobility/reducibility of manganese oxide. The oxidation of toluene was studied recently by Djinović et al. [122], who had examined the performance of monometallic CuO and bimetallic Cu-FeOx composites supported on KIL-2 silica. The utilization of two reducible oxides allowed increasing the amount and reactivity of oxygen species, which included adsorbed (O and O2) and lattice (O2−) oxygens at the Cu-FeOx interface, providing a substantial decrease of T90 that was ≈350 °C for the bimetallic cluster instead of ≈450 °C of monometallic copper oxide, whereas the FeOx/KIL-2 silica reached only 30% of toluene conversion at 450 °C.
Abdullah et al. [123] investigated the oxidation of a Cl-VOC mixture (dichloromethane (DCM), trichloroethylene (TCE) and trichloromethane (TCM)) with Cu-Cr/ZMS-5. Interestingly, in this case, the presence of water vapor in the gas feed enhanced the total oxidation to CO2. The presence of water vapor favored the formation of reactive carbocations. Furthermore, H2O was beneficial in blocking chlorine-transfer reactions. Indeed, an important deactivation effect was found with the bimetallic catalyst at a higher Cl/H gas feed ratio, and chlorination led to a decrease in metals’ reducibility that resulted in a low degradation efficiency. The reaction was driven by an MvK mechanism.
At this point, it is possible to highlight some differences through comparison of the catalytic performance of gold-based bimetallic samples with the others reported above. For the nongold-containing samples category, supports with a high surface area or with a tunable pore size distribution (silica, zeolite, alumina, etc.) were preferred to favor the dispersion of the active metals. Platinum-based samples gave the best results, with a second metal that in many cases enhances catalytic stability rather than overall VOC conversion. Although noble-metals–bimetallic catalysts showed the best performance, in recent years, in order to reduce the high cost of these catalysts, the addition or replacement of at least one of the noble metals with a cheaper transition metal is an interesting approach to reduce the total material cost while maintaining an acceptable catalytic activity.

3. Bimetallic Catalysts for the Photocatalytic Oxidation of VOCs

The urgent request for a “greener” and sustainable industrial chemistry has driven a huge field of research towards alternative ways to treat VOCs instead of catalytic combustion. Among the various AOPs (see the Introduction section), photocatalytic oxidation is the most applied process. With respect to catalytic thermal oxidation, this technique allows working at room temperature, exploiting the chemicophysical processes activated by an appropriate light radiation interacting with the surface of a semiconductor photocatalyst [40,124,125,126]. Specifically, dangerous organic compounds are oxidized by hydroxyl, and super oxide radicals are generated by the interaction between the photoelectrons and photoholes of the photocatalyst with water and oxygen. These photoelectrons and photoholes are formed when an adequate wavelength (λ ≤ of the band gap, Eg, of the semiconductor) irradiates the photocatalyst [127,128]. The most used photocatalysts are metal oxides or sulfides such as TiO2, ZnO, WO3, ZrO2, CeO2, Fe2O3, ZnS, and CdS, and among them, TiO2 and ZnO are the most used [129,130]. Due to its properties, such as its nontoxicity, relatively low cost, and high activity, especially under UV irradiation, titanium dioxide was deeply investigated both in academic and industrial research [124,126]. With a band gap varying from 3.0 to 3.2 eV depending on the crystalline form, TiO2 is able to exploit only 5% of solar radiation, thus limiting its practical applications. Photothermocatalytic oxidation is a multicatalytic approach that accepts the contemporaneous utilization of a light source to activate the photocatalyst, and thermal heat to boost the conversion of organic molecules and to increase the yield to CO2. A proper structural and/or chemical modification of titanium dioxide together with this multicatalytic approach can be considered a suitable solution to decrease the total energy consumption, maintaining the high conversion values typical of thermocatalytic oxidation of VOCs [131,132,133]. Another connected strategy to increase the photocatalytic performance of titanium dioxide under solar/visible light irradiation is doping with metal or nonmetal elements [134,135,136], and a not yet largely explored strategy is the combination of a bimetallic alloy with TiO2 [137,138]. The same metals typically employed for catalytic oxidation of VOCs, such as Au, Ag, Pt, Pd, and Cu in the nanoparticle size, if irradiated with (usually) a visible light irradiation, allow exploiting the localized surface plasmon resonance (LSPR) through collective oscillations of the electrons in the surface of the metal nanoparticles. This effect, combined with the photocatalytic properties of TiO2, is a performance solution to obtain a visible-light-driven photocatalyst [124,139,140]. In addition, for this particular application, bimetallic compounds can help to overcome some of the drawbacks of single metals. For example, the LSPR of some noble metals such as Pd, Pt, and Rh is not efficiently activated by solar irradiation, and a possible combination with the most effective plasmonic metals, such as Au, Ag or Cu, leads to takeing advantages of both the LSPR effect and of the reactive catalytic behavior of the other noble metals [137].
For the above considerations, in VOC photo-oxidation, the most investigated system was the Au-Pd bimetallic compound joined with a semiconductor photocatalyst [141,142,143,144,145]. In these materials, the good affinity of gold and palladium, already discussed in terms of thermocatalytic performance (Section 2.1.1.), is in this case utilized to increase the photocatalytic performances of titanium dioxide or of another semiconductor oxide.
Colmenares et al. [141] synthetized an Au-Pd/TiO2 photocatalyst with the original technique of sonophotodeposition (Figure 9). The bimetallic sample exhibited high activity (83%) and good selectivity (70%) in the partial oxidation of methanol to methyl formate after 120 min of UV irradiation (125 W mercury lamp λmax = 365 nm). Although the bimetallic catalyst showed a low selectivity to CO2 (≈30%), demonstrating that this approach is better suitable for selective oxidation than the total oxidation of VOCs, the reported material synthesis and adopted reaction conditions are a fascinating way to obtain results with an energy-efficient procedure and a selective photocatalyst in a short time and under mild conditions.
In a further study, the same research group [142] developed a density functional methodology to analyze the reaction mechanism of the selective photo-oxidation of methanol on the bimetallic Au-Pd/TiO2 sample. The theoretical investigation showed, as with the formation of a synergistic interaction between gold and palladium, a superior photoelectron–hole separation, was verified in comparison with the monometallic samples. Furthermore, it was shown that to favor total photo-oxidation to CO2, the dissociation of molecular oxygen should be driven preferentially on Pd to favor the formation of PdO sites, where complete oxidation (no methyl formate formation) to carbon dioxide occurred.
Cybula et al. [143] investigated the performance of an Au-Pd bimetallic sample supported on rutile TiO2 synthetized with a water in oil microemulsion methodology, in the photocatalytic oxidation of toluene and phenol under visible light irradiation (25 LEDs (λmax = 415 nm)). In particular, the authors focused on the effect of calcination temperature on materials’ preparation. The bimetallic sample calcined at 350 °C achieved 65% of toluene degradation and 22% of phenol conversion after 60 min of visible light irradiation. The performances were inferior compared to the photoactivity of monometallic palladium (79% in the toluene degradation and 24% in the phenol removal); however, the synergistic effect combined with a strong metals–support interaction was better exploited in the UV-vis tests, where the intrinsic photoactivity of rutile TiO2 also made a substantial contribution in removal efficiency. In fact, with the bimetallic sample, 100% of phenol degradation was achieved after 60 min of irradiation instead of the 56% of Pd/TiO2.
The interaction of the gold–palladium compound with other semiconductors was examined by the research group of Zhang et al. [144,145]. The photocatalytic oxidation of the trichloroethylene was studied on Au-Pd/BiPO4 nanorods and on Au-Pd/MoO3 nanowires. Interestingly, with the deposition of the Au-Pd alloy on the surface of the BiPO4 nanorod, the photocatalytic degradation rate increased quickly, being about 25 times higher compared to that achieved with bare BiPO4. The authors proposed, on the basis of the characterization measurements, the reaction mechanism illustrated in Figure 10. Under visible light irradiation (solar simulator with a 440 nm cut-off filter), Au-Pd/BiPO4 were excited due to the LSPR of the Au-Pd alloy. An effective charge carrier separation was achieved due to electron transfer from the conduction band (CB) of BiPO4 to the Au-Pd surface interface, whereas the photoholes remained confined in the valence band (VB) of BiPO4. Subsequently, the same photoelectrons present in the surface of the Au-Pd alloy reacted with the oxygens in the gas-phase that were successively reduced into superoxide radicals. These radicals together with the holes in the VB of BiPO4 oxidize the trichloroethylene in water and carbon dioxide.
The synergistic effect between gold and palladium nanoparticles that permitted increasing charge carrier separation, enhancing in this way photocatalytic activity, was verified, also employing MoO3 nanowires as a semiconductor photocatalyst [145].
The exploration of the synergistic effect with a bimetallic alloy in the photocatalytic oxidation of VOCs has not yet been largely investigated in the literature, and only fa ew examples are present considering other compounds instead of the Au-Pd system [146,147,148,149]. The possibility of favoring a selective oxidation route due to the high selectivity of gold–silver nanoparticles was examined by Han et al. [146] even in partial oxidation of methanol to methyl formate. In this case, the Au-Ag/TiO2 powders prepared via chemical reduction showed good results with a methanol conversion of 90% and a selectivity to methyl formate of about 85% under UV irradiation (500 W high pressure mercury lamp, λmax = 365 nm). Similarly to the Au-Pd systems, the bimetallic alloy enhanced the photoelectron–photohole separation with electron transfer from the conduction band of TiO2 (excited by the UV irradiation) to the gold–silver surface interface.
Another alloy with silver, i.e., Ag-Pt, was studied by Zieli’nska-Jurek et al. [147] in the photo-oxidation of toluene under visible light irradiation (LEDs, λmax = 415 nm). The authors found an interesting correlation regarding the order of deposition of the Ag-Pt/TiO2 photocatalysts prepared by sol–gel. In particular, the best sample (best photoactivity) was the material where the silver precursor was added before the platinum one. It was fundamental, in fact, to obtain particles with a definite size and dispersion (Ag-Pt size between 6–12 nm). In this way, it was possible to increase the toluene degradation rate with respect to the monometallic samples. By contrast, the bimetallic sample prepared with a simultaneous addition of metals precursors on TiO2 gave a lower photoactivity and different metal size and distribution. The authors concluded that platinum size had a greater influence than silver in determining overall photocatalytic activity. Recently, the same research group evaluated photocatalytic performance in both toluene and acetaldehyde degradation and of Penicillium chrysogenum, a dangerous fungus present in the indoor environment with Ag-Pt/TiO2 and Cu-Pt/TiO2 samples [148]. Both bimetallic samples showed a higher fungicidal activity under visible light irradiation than bare TiO2, whereas in VOC degradation, the Ag-Pt system was better-performing compared to Cu-Pt. The peculiar activity of both bimetallic samples was ascribed to the interfacial charge transfer process between the two metals and the TiO2 confirmed by the quenching of fluorescence (i.e., intensity diminution of the TiO2 photolumiscence bands) due to the presence of the metal alloy.
Wolski et al. [149] studied the mechanism of methanol photo-oxidation on bimetallic Au-Cu catalyst supported on Nb2O5 with an in operando IR methodology under both UV and visible light irradiation. Interestingly, they found that photocatalytic activity is strictly related to the light sources used and to the number of Brønsted/Lewis acid sites present on the surface of the catalysts. Specifically, under visible light irradiation, the synergism between gold and copper led to an increase in the amount of Brønsted/Lewis acid sites on the niobia, with a consequent higher activity of bimetallic samples compared to that of monometallic and pure Nb2O5 samples. Furthermore, the total oxidation to CO2 was favored. By contrast, with UV light irradiation, the major activation of niobia (Eg ≈ 3.2 eV) favored selective oxidation into dimethoxymethane, formaldehyde, and methyl formate.
In this short chapter, the state-of-the-art of the application of bimetallic structures as chemical modifiers of conventional and unconventional semiconductor photocatalysts was examined. This approach is relatively new, and the effects of alloy synergism on the photocatalytic process are currently under investigation. The promising results, especially obtained by combining the LSPR effect of both noble and transition metals with semiconductor photoactivity, together with a possible multicatalytic strategy (i.e., a photothermo atalytic approach employing a bimetallic/semiconductor catalyst and a solar/visible light source) could in the future be a fascinating strategy to develop a greener and sustainable technology applied to the removal of volatile organic compounds.

4. Conclusions

In this review, the application of bimetallic catalysts for VOC oxidation was examined in terms of catalytic activity and physicochemical properties. Among the various systems, gold-based bimetallic catalysts exhibited a good performance in the degradation of a wide range of VOCs. The presence of nanosized gold was essential to decreasing the light-off temperature of VOC oxidation, whereas interaction with the second metal allowed increasing the reactivity of the employed support or enhancing oxygen activation. Although platinum-based bimetallic samples usually did not overcome the degradation yields achieved with the monometallic platinum catalysts, they showed a substantial improvement of catalytic stability due to the synergistic effect between platinum and the second noble or transition metal. Finally, the application of the already stated synergisms in catalytic thermo-oxidation, for example, strong Au-Pd interaction, can be successfully transferred to new technologies for VOC abatement, such as photocatalytic oxidation, with the exploitation of new mutual effects such as surface plasmon resonance combined with the high reactivity of noble/transition metals. This can be a promising strategy to achieve significant progress in the technologies applied to the improvement of air quality.

Funding

R.F. thanks the PON project “AIM” founded by the European Social Found (ESF) CUP: E66C18001220005 for the financial support.

Acknowledgments

This review was written during the Italian quarantine due to COVID-19. I want to dedicate this work to all the people who could not stay at home because they fought against the virus. R.F. thanks S. Scirè and C. Crisafulli who have positively transferred the knowledge and the passion for the research in the field of heterogeneous catalysis focused on environmental applications.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Yang, C.; Miao, G.; Pi, Y.; Xia, Q.; Wu, J.; Li, Z.; Xiao, J. Abatement of various types of VOCs by adsorption/catalytic oxidation: A review. Chem. Eng. J. 2019, 370, 1128–1153. [Google Scholar] [CrossRef]
  2. He, C.; Cheng, J.; Zhang, X.; Douthwaite, M.; Pattisson, S.; Hao, Z. Recent advances in the catalytic oxidation of volatile organic compounds: A review based on pollutant sorts and sources. Chem. Rev. 2019, 119, 4471–4568. [Google Scholar] [CrossRef] [PubMed]
  3. Amann, M.; Lutz, M. The revision of the air quality legislation in the European Union related to ground-level ozone. J. Hazard. Mater. 2000, 78, 41–62. [Google Scholar] [CrossRef]
  4. Dumanoglu, Y.; Kara, M.; Altiok, H.; Odabasi, M.; Elbir, T.; Bayram, A. Spatial and seasonal variation and source apportionment of volatile organic compounds (VOCs) in a heavily industrialized region. Atmos. Environ. 2014, 98, 168–178. [Google Scholar] [CrossRef]
  5. Hu, R.; Liu, G.; Zhang, H.; Xue, H.; Wang, X.; Lam, P.K.S. Odor pollution due to industrial emission of volatile organic compounds: A case study in Hefei, China. J. Clean. Prod. 2020, 246, 119075. [Google Scholar] [CrossRef]
  6. Liao, H.T.; Chou, C.C.K.; Chow, J.C.; Watson, J.G.; Hopke, P.K.; Wu, C.F. Source and risk apportionment of selected VOCs and PM2.5 species using partially constrained receptor models with multiple time resolution data. Environ. Pollut. 2015, 205, 121–130. [Google Scholar] [CrossRef] [PubMed]
  7. Dhamodharan, K.; Varma, V.S.; Veluchamy, C.; Pugazhendhi, A.; Rajendran, K. Emission of volatile organic compounds from composting: A review on assessment, treatment and perspectives. Sci. Total Environ. 2019, 695, 133725. [Google Scholar] [CrossRef]
  8. Tørseth, K.; Aas, W.; Breivik, K.; Fjeraa, A.M.; Fiebig, M.; Hjellbrekke, A.G.; Lund Myhre, C.; Solberg, S.; Yttri, K.E. Introduction to the European Monitoring and Evaluation Programme (EMEP) and observed atmospheric composition change during 1972–2009. Atmos. Chem. Phys. 2012, 12, 5447–5481. [Google Scholar] [CrossRef] [Green Version]
  9. Huang, B.; Lei, C.; Wei, C.; Zeng, G. Chlorinated volatile organic compounds (Cl-VOCs) in environment—Sources, potential human health impacts, and current remediation technologies. Environ. Int. 2014, 71, 118–138. [Google Scholar] [CrossRef]
  10. Kumar, V.; Lee, Y.S.; Shin, J.W.; Kim, K.H.; Kukkar, D.; Tsang, Y.F. Potential applications of graphene-based nanomaterials as adsorbent for removal of volatile organic compounds. Environ. Int. 2020, 135, 105356. [Google Scholar] [CrossRef]
  11. Parmar, G.R.; Rao, N.N. Emerging control technologies for volatile organic compounds. Crit. Rev. Environ. Sci. Technol. 2008, 39, 41–78. [Google Scholar] [CrossRef]
  12. Zhu, L.; Shen, D.; Luo, K.H. A critical review on VOCs adsorption by different porous materials: Species, mechanisms and modification methods. J. Hazard. Mater. 2020, 389, 122102. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, X.; Gao, B.; Creamer, A.E.; Cao, C.; Li, Y. Adsorption of VOCs onto engineered carbon materials: A review. J. Hazard. Mater. 2017, 338, 102–123. [Google Scholar] [CrossRef] [PubMed]
  14. 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]
  15. Spivey, J.J. Complete catalytic oxidation of volatile organics. Ind. Eng. Chem. Res. 1987, 26, 2165–2180. [Google Scholar] [CrossRef]
  16. Liotta, L.F. Catalytic oxidation of volatile organic compounds on supported noble metals. Appl. Catal. B Environ. 2010, 100, 403–412. [Google Scholar] [CrossRef]
  17. Zhang, Z.; Jiang, Z.; Shangguan, W. Low-temperature catalysis for VOCs removal in technology and application: A state-of-the-art review. Catal. Today 2016, 264, 270–278. [Google Scholar] [CrossRef]
  18. Huang, H.; Xu, Y.; Feng, Q.; Leung, D.Y.C. Low temperature catalytic oxidation of volatile organic compounds: A review. Catal. Sci. Technol. 2015, 5, 2649–2669. [Google Scholar] [CrossRef]
  19. Fiorenza, R.; Balsamo, S.A.; D’Urso, L.; Scirè, S.; Brundo, M.V.; Pecoraro, R.; Scalisi, E.M.; Privitera, V.; Impellizzeri, G. CeO2 for Water Remediation: Comparison of Various Advanced Oxidation Processes. Catalysts 2020, 10, 446. [Google Scholar] [CrossRef] [Green Version]
  20. Tokumura, M.; Nakajima, R.; Znad, H.T.; Kawase, Y. Chemical absorption process for degradation of VOC gas using heterogeneous gas-liquid photocatalytic oxidation: Toluene degradation by photo-Fenton reaction. Chemosphere 2008, 73, 768–775. [Google Scholar] [CrossRef]
  21. Malakar, S.; Das, P.S.; Baskaran, D.; Rajamanickam, R. Comparative study of biofiltration process for treatment of VOCs emission from petroleum refinery wastewater—A review. Environ. Technol. Innov. 2017, 8, 441–461. [Google Scholar] [CrossRef]
  22. Pettit, T.; Irga, P.J.; Torpy, F.R. Towards practical indoor air phytoremediation: A review. Chemosphere 2018, 208, 960–974. [Google Scholar] [CrossRef] [PubMed]
  23. Jakubek, T.; Hudy, C.; Indyka, P.; Nowicka, E.; Golunski, S.; Kotarba, A. Effect of noble metal addition to alkali-exchanged cryptomelane on the simultaneous soot and VOC combustion activity. Catal. Commun. 2019, 132, 105807. [Google Scholar] [CrossRef]
  24. Gallastegi-Villa, M.; Aranzabal, A.; Romero-Sáez, M.; González-Marcos, J.A.; González-Velasco, J.R. Catalytic activity of regenerated catalyst after the oxidation of 1,2-dichloroethane and trichloroethylene. Chem. Eng. J. 2014, 241, 200–206. [Google Scholar] [CrossRef]
  25. De Rivas, B.; Sampedro, C.; García-Real, M.; López-Fonseca, R.; Gutiérrez-Ortiz, J.I. Promoted activity of sulphated Ce/Zr mixed oxides for chlorinated VOC oxidative abatement. Appl. Catal. B Environ. 2013, 129, 225–235. [Google Scholar] [CrossRef]
  26. Barakat, T.; Rooke, J.C.; Genty, E.; Cousin, R.; Siffert, S.; Su, B.L. Gold catalysts in environmental remediation and water-gas shift technologies. Energy Environ. Sci. 2013, 6, 371–391. [Google Scholar] [CrossRef]
  27. Matějová, L.; Topka, P.; Kaluža, L.; Pitkäaho, S.; Ojala, S.; Gaálová, J.; Keiski, R.L. Total oxidation of dichloromethane and ethanol over ceria-zirconia mixed oxide supported platinum and gold catalysts. Appl. Catal. B Environ. 2013, 142, 54–64. [Google Scholar] [CrossRef]
  28. Guo, Y.; Yang, D.P.; Liu, M.; Zhang, X.; Chen, Y.; Huang, J.; Li, Q.; Luque, R. Enhanced catalytic benzene oxidation over a novel waste-derived Ag/eggshell catalyst. J. Mater. Chem. A 2019, 7, 8832–8844. [Google Scholar] [CrossRef]
  29. Gaálová, J.; Topka, P.; Kaluža, L.; Soukup, K.; Barbier, J. Effect of gold loading on ceria-zirconia support in total oxidation of VOCs. Catal. Today 2019, 333, 190–195. [Google Scholar] [CrossRef]
  30. Radic, N.; Grbic, B.; Terlecki-Baricevic, A. Kinetics of deep oxidation of n-hexane and toluene over Pt/Al 2 O 3 catalysts: Platinum crystallite size effect. Appl. Catal. B Environ. 2004, 50, 153–159. [Google Scholar] [CrossRef]
  31. Topka, P.; Dvořáková, M.; Kšírová, P.; Perekrestov, R.; Čada, M.; Balabánová, J.; Koštejn, M.; Jirátová, K.; Kovanda, F. Structured cobalt oxide catalysts for VOC abatement: The effect of preparation method. Environ. Sci. Pollut. Res. 2019, 27, 7608–7617. [Google Scholar] [CrossRef]
  32. Rodríguez, M.L.; Cadús, L.E.; Borio, D.O. VOCs abatement in adiabatic monolithic reactors: Heat effects, transport limitations and design considerations. Chem. Eng. J. 2016, 306, 86–98. [Google Scholar] [CrossRef]
  33. Liu, B.; Zhan, Y.; Xie, R.; Huang, H.; Li, K.; Zeng, Y.; Shrestha, R.P.; Kim Oanh, N.T.; Winijkul, E. Efficient photocatalytic oxidation of gaseous toluene in a bubbling reactor of water. Chemosphere 2019, 233, 754–761. [Google Scholar] [CrossRef] [PubMed]
  34. Perez, V.; Miachon, S.; Dalmon, J.A.; Bredesen, R.; Pettersen, G.; Rader, H.; Simon, C. Preparation and characterisation of a Pt/ceramic catalytic membrane. Sep. Purif. Technol. 2001, 25, 33–38. [Google Scholar] [CrossRef]
  35. Trovarelli, A.; Llorca, J. Ceria catalysts at nanoscale: How do crystal shapes shape catalysis? ACS Catal. 2017, 7, 4716–4735. [Google Scholar] [CrossRef]
  36. Li, J.; Liu, H.; Deng, Y.; Liu, G.; Chen, Y.; Yang, J. Emerging nanostructured materials for the catalytic removal of volatile organic compounds. Nanotechnol. Rev. 2016, 5, 147–181. [Google Scholar] [CrossRef]
  37. Busca, G.; Berardinelli, S.; Resini, C.; Arrighi, L. Technologies for the removal of phenol from fluid streams: A short review of recent developments. J. Hazard. Mater. 2008, 160, 265–288. [Google Scholar] [CrossRef]
  38. Van Durme, J.; Dewulf, J.; Leys, C.; van Langenhove, H. Combining non-thermal plasma with heterogeneous catalysis in waste gas treatment: A review. Appl. Catal. B Environ. 2008, 78, 324–333. [Google Scholar] [CrossRef] [Green Version]
  39. Veerapandian, S.K.P.; Leys, C.; De Geyter, N.; Moren, R. Abatement of VOCs using packed bed non-thermal plasma reactors: A review. Catalysts 2017, 7, 113. [Google Scholar] [CrossRef]
  40. Mamaghani, A.H.; Haghighat, F.; Lee, C.S. Photocatalytic oxidation technology for indoor environment air purification: The state-of-the-art. Appl. Catal. B Environ. 2017, 203, 247–269. [Google Scholar] [CrossRef]
  41. Stytsenko, V.D. Surface modified bimetallic catalysts: Preparation, characterization, and applications. Appl. Catal. A Gen. 1995, 126, 1–26. [Google Scholar] [CrossRef]
  42. Rodriguez, J. Physical and chemical properties of bimetallic surfaces. Surf. Sci. Rep. 1996, 24, 223–287. [Google Scholar] [CrossRef]
  43. Singh, A.K.; Xu, Q. Synergistic catalysis over bimetallic alloy nanoparticles. ChemCatChem 2013, 5, 652–676. [Google Scholar] [CrossRef]
  44. Sankar, M.; Dimitratos, N.; Miedziak, P.J.; Wells, P.P.; Kiely, C.J.; Hutchings, G.J. Designing bimetallic catalysts for a green and sustainable future. Chem. Soc. Rev. 2012, 41, 8099–8139. [Google Scholar] [CrossRef] [PubMed]
  45. Sinfelt, J.H. Supported “bimetallic cluster” catalysts. J. Catal. 1973, 29, 308–315. [Google Scholar] [CrossRef]
  46. Blaser, H.-U.; Malan, C.; Pugin, B.; Spindler, F.; Steiner, H.; Studer, M. Selective hydrogenation for fine chemicals: Recent trends and new developments. ChemInform 2003, 345, 103–151. [Google Scholar] [CrossRef]
  47. Hu, M.; Jin, L.; Zhu, Y.; Zhang, L.; Lu, X.; Kerns, P.; Su, X.; Cao, S.; Gao, P.; Suib, S.L.; et al. Self-limiting growth of ligand-free ultrasmall bimetallic nanoparticles on carbon through under temperature reduction for highly efficient methanol electrooxidation and selective hydrogenation. Appl. Catal. B Environ. 2020, 264, 118553. [Google Scholar] [CrossRef]
  48. Zhang, J.; Wang, H.; Dalai, A.K. Development of stable bimetallic catalysts for carbon dioxide reforming of methane. J. Catal. 2007, 249, 300–310. [Google Scholar] [CrossRef]
  49. Li, L.; Zuo, S.; An, P.; Wu, H.; Hou, F.; Li, G.; Liu, G. Hydrogen production via steam reforming of n-dodecane over NiPt alloy catalysts. Fuel 2020, 262, 116469. [Google Scholar] [CrossRef]
  50. Fiorenza, R.; Crisafulli, C.; Scirè, S. H2 purification through preferential oxidation of CO over ceria supported bimetallic Au-based catalysts. Int. J. Hydrogen Energy 2016, 41, 19390–19398. [Google Scholar] [CrossRef]
  51. Fiorenza, R.; Spitaleri, L.; Gulino, A.; Scirè, S. Ru–Pd bimetallic catalysts supported on CeO2-MnOx oxides as efficient systems for H2 purification through CO preferential Oxidation. Catalysts 2018, 8, 203. [Google Scholar] [CrossRef] [Green Version]
  52. Fiorenza, R.; Scirè, S.; Venezia, A.M. Carbon supported bimetallic Ru-Co catalysts for H2 production through NaBH4 and NH3BH3 hydrolysis. Int. J. Energy Res. 2018, 42, 1183–1195. [Google Scholar] [CrossRef]
  53. Fiorenza, R.; Spitaleri, L.; Gulino, A.; Sciré, S. High-performing Au-Ag bimetallic catalysts supported on macro-mesoporous CeO2 for preferential oxidation of CO in H2-rich gases. Catalysts 2020, 10, 49. [Google Scholar] [CrossRef] [Green Version]
  54. Luo, L.; Chen, S.; Xu, Q.; He, Y.; Dong, Z.; Zhang, L.; Zhu, J.; Du, Y.; Yang, B.; Wang, C. Dynamic atom clusters on AuCu nanoparticle surface during CO oxidation. J. Am. Chem. Soc. 2020, 142, 4022–4027. [Google Scholar] [CrossRef]
  55. Guo, S.; Dong, S.; Wang, E. Three-dimensional Pt-on-Pd bimetallic nanodendrites supported on graphene nanosheet: Facile synthesis and used as an advanced nanoelectrocatalyst for methanol oxidation. ACS Nano 2010, 4, 547–555. [Google Scholar] [CrossRef] [PubMed]
  56. Zhang, H.; Jin, M.; Xia, Y. Enhancing the catalytic and electrocatalytic properties of Pt-based catalysts by forming bimetallic nanocrystals with Pd. Chem. Soc. Rev. 2012, 41, 8035–8049. [Google Scholar] [CrossRef]
  57. Nasrabadi, H.T.; Abbasi, E.; Davaran, S.; Kouhi, M.; Akbarzadeh, A. Bimetallic nanoparticles: Preparation, properties, and biomedical applications. Artif. Cells Nanomed. Biotechnol. 2016, 44, 376–380. [Google Scholar] [CrossRef]
  58. Ferrando, R.; Jellinek, J.; Johnston, R.L. Nanoalloys: From theory to applications of alloy clusters and nanoparticles. Chem. Rev. 2008, 108, 845–910. [Google Scholar] [CrossRef] [PubMed]
  59. Duan, S.; Wang, R. Bimetallic nanostructures with magnetic and noble metals and their physicochemical applications. Prog. Nat. Sci. Mater. Int. 2013, 23, 113–126. [Google Scholar] [CrossRef] [Green Version]
  60. Dehghan Banadaki, A.; Kajbafvala, A. Recent advances in facile synthesis of bimetallic nanostructures: An overview. J. Nanomater. 2014, 2014, 1–28. [Google Scholar] [CrossRef] [Green Version]
  61. Duan, M.; Jiang, L.; Zeng, G.; Wang, D.; Tang, W.; Liang, J.; Wang, H.; He, D.; Liu, Z.; Tang, L. Bimetallic nanoparticles/metal-organic frameworks: Synthesis, applications and challenges. Appl. Mater. Today 2020, 19, 100564. [Google Scholar] [CrossRef]
  62. Redina, E.A.; Kirichenko, O.A.; Greish, A.A.; Kucherov, A.V.; Tkachenko, O.P.; Kapustin, G.I.; Mishin, I.V.; Kustov, L.M. Preparation of bimetallic gold catalysts by redox reaction on oxide-supported metals for green chemistry applications. Catal. Today 2015, 246, 216–231. [Google Scholar] [CrossRef]
  63. Alexeev, O.S.; Gates, B.C. Supported bimetallic cluster catalysts. Ind. Eng. Chem. Res. 2003, 42, 1571–1587. [Google Scholar] [CrossRef]
  64. Bariås, O.A.; Holmen, A.; Blekkan, E.A. Propane dehydrogenation over supported Pt and Pt-Sn catalysts: Catalyst preparation, characterization, and activity measurements. J. Catal. 1996, 158, 1–12. [Google Scholar] [CrossRef]
  65. Zhou, S.; Kang, L.; Zhou, X.; Xu, Z.; Zhu, M. Pure acetylene semihydrogenation over Ni–Cu bimetallic catalysts: Effect of the Cu/Ni ratio on catalytic performance. Nanomaterials 2020, 10, 509. [Google Scholar] [CrossRef] [Green Version]
  66. Luisetto, I.; Tuti, S.; Di Bartolomeo, E. Co and Ni supported on CeO2 as selective bimetallic catalyst for dry reforming of methane. Int. J. Hydrogen Energy 2012, 37, 15992–15999. [Google Scholar] [CrossRef]
  67. Aguirre, A.; Zanella, R.; Barrios, C.; Hernández, S.; Bonivardi, A.; Collins, S.E. Gold stabilized with iridium on ceria–niobia catalyst: Activity and stability for CO oxidation. Top. Catal. 2019, 62, 977–988. [Google Scholar] [CrossRef]
  68. 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]
  69. Xia, S.; Yuan, Z.; Wang, L.; Chen, P.; Hou, Z. Hydrogenolysis of glycerol on bimetallic Pd-Cu/solid-base catalysts prepared via layered double hydroxides precursors. Appl. Catal. A Gen. 2011, 403, 173–182. [Google Scholar] [CrossRef]
  70. Bönnemann, H.; Braun, G.; Brijoux, W.; Brinkmann, R.; Tilling, A.S.; Seevogel, K.; Siepen, K. Nanoscale colloidal metals and alloys stabilized by solvents and surfactants: Preparation and use as catalyst precursors. J. Organomet. Chem. 1996, 520, 143–162. [Google Scholar] [CrossRef]
  71. Minicò, S.; Scirè, S.; Crisafulli, C.; Galvagno, S. Influence of catalyst pretreatments on volatile organic compounds oxidation over gold/iron oxide. Appl. Catal. B Environ. 2001, 34, 277–285. [Google Scholar] [CrossRef]
  72. Scirè, S.; Liotta, L.F. Supported gold catalysts for the total oxidation of volatile organic compounds. Appl. Catal. B Environ. 2012, 125, 222–246. [Google Scholar] [CrossRef]
  73. Haruta, M. Gold as a novel catalyst in the 21st century: Preparation, working mechanism and applications. Gold Bull. 2004, 37, 27–36. [Google Scholar] [CrossRef] [Green Version]
  74. Haruta, M.; Yamada, N.; Kobayashi, T.; Iijima, S. Gold catalysts prepared by coprecipitation for low-temperature oxidation of hydrogen and of carbon monoxide. J. Catal. 1989, 115, 301–309. [Google Scholar] [CrossRef]
  75. Carrettin, S.; McMorn, P.; Johnston, P.; Griffin, K.; Hutchings, G.J. Selective oxidation of glycerol to glyceric acid using a gold catalyst in aqueous sodium hydroxide. Chem. Commun. 2002, 7, 696–697. [Google Scholar] [CrossRef]
  76. Prati, L.; Villa, A.; Jouve, A.; Beck, A.; Evangelisti, C.; Savara, A. Gold as a modifier of metal nanoparticles: Effect on structure and catalysis. Faraday Discuss. 2018, 208, 395–407. [Google Scholar] [CrossRef]
  77. Fajín, J.L.C.; Cordeiro, M.N.D.S.; Gomes, J.R.B. Catalytic reactions on model gold surfaces: Effect of surface steps and of surface doping. Catalysts 2011, 1, 40–51. [Google Scholar] [CrossRef] [Green Version]
  78. Hashmi, A.S.K.; Rudolph, M. Gold catalysis in total synthesis. Chem. Soc. Rev. 2008, 37, 1766–1775. [Google Scholar] [CrossRef]
  79. Liu, X.Y.; Wang, A.; Zhang, T.; Mou, C.Y. Catalysis by gold: New insights into the support effect. Nano Today 2013, 8, 403–416. [Google Scholar] [CrossRef]
  80. Xu, Q.; Lei, W.; Li, X.; Qi, X.; Yu, J.; Liu, G.; Wang, J.; Zhang, P. Efficient removal of formaldehyde by nanosized gold on well-defined CeO2 nanorods at room temperature. Environ. Sci. Technol. 2014, 48, 9702–9708. [Google Scholar] [CrossRef]
  81. Bell, A.T. The impact of nanoscience on heterogeneous catalysis. Science 2003, 299, 1688–1691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Daté, M.; Okumura, M.; Tsubota, S.; Haruta, M. Vital role of moisture in the catalytic activity of supported gold nanoparticles. Angew. Chem. Int. Ed. 2004, 43, 2129–2132. [Google Scholar] [CrossRef] [PubMed]
  83. Villa, A.; Wang, D.; Su, D.S.; Prati, L. New challenges in gold catalysis: Bimetallic systems. Catal. Sci. Technol. 2015, 5, 55–68. [Google Scholar] [CrossRef] [Green Version]
  84. Louis, C. Chemical preparation of supported bimetallic catalysts. Gold-based bimetallic, a case study. Catalysts 2016, 6, 110. [Google Scholar] [CrossRef]
  85. 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 Environ. 2012, 111, 218–224. [Google Scholar] [CrossRef]
  86. Barakat, T.; Rooke, J.C.; Chlala, D.; Cousin, R.; Lamonier, J.F.; Giraudon, J.M.; Casale, S.; Massiani, P.; Su, B.L.; Siffert, S. Oscillatory behavior of Pd-Au catalysts in toluene total oxidation. Catalysts 2018, 8, 574. [Google Scholar] [CrossRef] [Green Version]
  87. Xia, Y.; Xia, L.; Liu, Y.; Yang, T.; Deng, J.; Dai, H. Concurrent catalytic removal of typical volatile organic compound mixtures over Au-Pd/α-MnO2 nanotubes. J. Environ. Sci. (China) 2018, 64, 276–288. [Google Scholar] [CrossRef]
  88. Xie, S.; Deng, J.; Zang, S.; Yang, H.; Guo, G.; Arandiyan, H.; Dai, H. Au-Pd/3DOM Co3O4: Highly active and stable nanocatalysts for toluene oxidation. J. Catal. 2015, 322, 38–48. [Google Scholar] [CrossRef]
  89. Tabakova, T.; Ilieva, L.; Petrova, P.; Venezia, A.M.; Avdeev, G.; Zanella, R.; Karakirova, Y. Complete benzene oxidation over mono and bimetallic au-pd catalysts supported on fe-modified ceria. Chem. Eng. J. 2015, 260, 133–141. [Google Scholar] [CrossRef]
  90. Li, X.; Feng, J.; Perdjon, M.; Oh, R.; Zhao, W.; Huang, X.; Liu, S. Investigations of supported Au-Pd nanoparticles on synthesized CeO2 with different morphologies and application in solvent-free benzyl alcohol oxidation. Appl. Surf. Sci. 2020, 505, 144473. [Google Scholar] [CrossRef]
  91. Xie, S.; Liu, Y.; Deng, J.; Zhao, X.; Yang, J.; Zhang, K.; Han, Z.; Arandiyan, H.; Dai, H. Effect of transition metal doping on the catalytic performance of Au–Pd/3DOM Mn2O3 for the oxidation of methane and o-xylene. Appl. Catal. B Environ. 2017, 206, 221–232. [Google Scholar] [CrossRef]
  92. Kucherov, A.V.; Tkachenko, O.P.; Kirichenko, O.A.; Kapustin, G.I.; Mishin, I.V.; Klementiev, K.V.; Ojala, S.; Kustov, L.M.; Keiski, R. Nanogold-containing catalysts for low-temperature removal of S-VOC from air. Top. Catal. 2009, 52, 351–358. [Google Scholar] [CrossRef]
  93. Nagy, G.; Benkó, T.; Borkó, L.; Csay, T.; Horváth, A.; Frey, K.; Beck, A. Bimetallic Au-Ag/SiO2 catalysts: Comparison in glucose, benzyl alcohol and CO oxidation reactions. React. Kinet. Mech. Catal. 2015, 115, 45–65. [Google Scholar] [CrossRef]
  94. Bracey, C.L.; Ellis, P.R.; Hutchings, G.J. Application of copper-gold alloys in catalysis: Current status and future perspectives. Chem. Soc. Rev. 2009, 38, 2231–2243. [Google Scholar] [CrossRef] [PubMed]
  95. Chimentão, R.J.; Medina, F.; Fierro, J.L.G.; Llorca, J.; Sueiras, J.E.; Cesteros, Y.; Salagre, P. Propene epoxidation by nitrous oxide over Au–Cu/TiO2 alloy catalysts. J. Mol. Catal. A Chem. 2007, 274, 159–168. [Google Scholar] [CrossRef]
  96. Nevanperä, T.K.; Ojala, S.; Laitinen, T.; Pitkäaho, S.; Saukko, S.; Keiski, R.L. Catalytic oxidation of dimethyl disulfide over bimetallic Cu–Au and Pt–Au catalysts supported on γ-Al2O3, CeO2, and CeO2–Al2O3. Catalysts 2019, 9, 603. [Google Scholar] [CrossRef] [Green Version]
  97. Kim, K.J.; Ahn, H.G. Complete oxidation of toluene over bimetallic Pt-Au catalysts supported on ZnO/Al2O3. Appl. Catal. B Environ. 2009, 91, 308–318. [Google Scholar] [CrossRef]
  98. Kim, K.J.; Boo, S.-I.; Ahn, H.G. Preparation and characterization of the bimetallic Pt-Au/ZnO/Al2O3 catalysts: Influence of Pt-Au molar ratio on the catalytic activity for toluene oxidation. J. Ind. Eng. Chem. 2009, 15, 92–97. [Google Scholar] [CrossRef]
  99. Sreethawong, T.; Sukjit, D.; Ouraipryvan, P.; Schwank, J.W.; Chavadej, S. Oxidation of oxygenated volatile organic compound over monometallic and bimetallic Ru-Au catalysts. Catal. Lett. 2010, 138, 160–170. [Google Scholar] [CrossRef]
  100. Calzada, L.A.; Collins, S.E.; Han, C.W.; Ortalan, V.; Zanella, R. Synergetic effect of bimetallic Au-Ru/TiO2 catalysts for complete oxidation of methanol. Appl. Catal. B Environ. 2017, 207, 79–92. [Google Scholar] [CrossRef]
  101. Torrente-Murciano, L.; Solsona, B.; Agouram, S.; Sanchis, R.; López, J.M.; García, T.; Zanella, R. Low temperature total oxidation of toluene by bimetallic Au-Ir catalysts. Catal. Sci. Technol. 2017, 7, 2886–2896. [Google Scholar] [CrossRef]
  102. Solsona, B.; Pérez-Cabero, M.; Vázquez, I.; Dejoz, A.; García, T.; Álvarez-Rodríguez, J.; El-Haskouri, J.; Beltrán, D.; Amorós, P. Total oxidation of VOCs on Au nanoparticles anchored on Co doped mesoporous UVM-7 silica. Chem. Eng. J. 2012, 187, 391–400. [Google Scholar] [CrossRef]
  103. Bonelli, R.; Lucarelli, C.; Pasini, T.; Liotta, L.F.; Zacchini, S.; Albonetti, S. Total oxidation of volatile organic compounds on Au/FeOx catalysts supported on mesoporous SBA-15 silica. Appl. Catal. A Gen. 2011, 400, 54–60. [Google Scholar] [CrossRef] [Green Version]
  104. Albonetti, S.; Bonelli, R.; Delaigle, R.; Gaigneaux, E.M.; Femoni, C.; Riccobene, P.M.; Scirè, S.; Tiozzo, C.; Zacchini, S.; Trifirò, F. Design of nano-sized FeOx and Au/FeOx catalysts for total oxidation of VOC and preferential oxidation of CO. In Studies in Surface Science and Catalysis; Elsevier: Amsterdam, The Netherlands, 2010; Volume 175, pp. 785–788. [Google Scholar]
  105. Fu, X.; Liu, Y.; Yao, W.; Wu, Z. One-step synthesis of bimetallic Pt-Pd/MCM-41 mesoporous materials with superior catalytic performance for toluene oxidation. Catal. Commun. 2016, 83, 22–26. [Google Scholar] [CrossRef]
  106. Kim, H.S.; Kim, T.W.; Koh, H.L.; Lee, S.H.; Min, B.R. Complete benzene oxidation over Pt-Pd bimental catalyst supported on γ-alumina: Influence of Pt-Pd ratio on the catalytic activity. Appl. Catal. A Gen. 2005, 280, 125–131. [Google Scholar] [CrossRef]
  107. Kim, H.S.; Min, M.K.; Song, M.W.; Park, J.W.; Min, B.R. XPS analysis of the effect of Pt addition to Pd catalysts for complete benzene oxidation. React. Kinet. Catal. Lett. 2004, 81, 251–257. [Google Scholar] [CrossRef]
  108. Sharma, R.K.; Zhou, B.; Tong, S.; Chuang, K.T. Catalytic destruction of volatile organic compounds using supported platinum and palladium hydrophobic catalysts. Ind. Eng. Chem. Res. 1995, 34, 4310–4317. [Google Scholar] [CrossRef]
  109. Liu, X.; Zeng, J.; Shi, W.; Wang, J.; Zhu, T.; Chen, Y. Catalytic oxidation of benzene over ruthenium-cobalt bimetallic catalysts and study of its mechanism. Catal. Sci. Technol. 2017, 7, 213–221. [Google Scholar] [CrossRef]
  110. Ye, M.; Chen, L.; Liu, X.; Xu, W.; Zhu, T.; Chen, G. Catalytic oxidation of chlorobenzene over ruthenium-ceria bimetallic catalysts. Catalysts 2018, 8, 116. [Google Scholar] [CrossRef] [Green Version]
  111. De La Peña O’Shea, V.A.; Álvarez-Galván, M.C.; Fierro, J.L.G.; Arias, P.L. Influence of feed composition on the activity of Mn and PdMn/Al 2O3 catalysts for combustion of formaldehyde/methanol. Appl. Catal. B Environ. 2005, 57, 191–199. [Google Scholar]
  112. Kim, S.C.; Park, Y.K.; Nah, J.W. Property of a highly active bimetallic catalyst based on a supported manganese oxide for the complete oxidation of toluene. Powder Technol. 2014, 266, 292–298. [Google Scholar] [CrossRef]
  113. Jodaei, A.; Salari, D.; Niaei, A.; Khatamian, M.; Çaylak, N. Preparation of Ag-M (M: Fe, Co and Mn)-ZSM-5 bimetal catalysts with high performance for catalytic oxidation of ethyl acetate. Environ. Technol. 2011, 32, 395–406. [Google Scholar] [CrossRef] [PubMed]
  114. Rintramee, K.; Föttinger, K.; Rupprechter, G.; Wittayakun, J. Ethanol adsorption and oxidation on bimetallic catalysts containing platinum and base metal oxide supported on MCM-41. Appl. Catal. B Environ. 2012, 115–116, 225–235. [Google Scholar] [CrossRef]
  115. Chantaravitoon, P.; Chavadej, S.; Schwank, J. Temperature-programmed desorption of methanol and oxidation of methanol on Pt-Sn/Al2O3 catalysts. Chem. Eng. J. 2004, 97, 161–171. [Google Scholar] [CrossRef]
  116. Baranowska, K.; Okal, J. Performance and stability of the Ru-Re/γ-Al2O3 catalyst in the total oxidation of propane: Influence of the order of impregnation. Catal. Lett. 2016, 146, 72–81. [Google Scholar] [CrossRef]
  117. Yue, L.; He, C.; Zhang, X.; Li, P.; Wang, Z.; Wang, H.; Hao, Z. Catalytic behavior and reaction routes of MEK oxidation over Pd/ZSM-5 and Pd-Ce/ZSM-5 catalysts. J. Hazard. Mater. 2013, 244, 613–620. [Google Scholar] [CrossRef]
  118. Jodaei, A.; Niaei, A.; Salari, D. Performance of nanostructure Fe-Ag-ZSM-5 catalysts for the catalytic oxidation of volatile organic compounds: Process optimization using response surface methodology. Korean J. Chem. Eng. 2011, 28, 1665–1671. [Google Scholar] [CrossRef]
  119. Izadkhah, B.; Nabavi, S.R.; Niaei, A.; Salari, D.; Mahmuodi Badiki, T.; Çaylak, N. Design and optimization of Bi-metallic Ag-ZSM5 catalysts for catalytic oxidation of volatile organic compounds. J. Ind. Eng. Chem. 2012, 18, 2083–2091. [Google Scholar] [CrossRef]
  120. Qu, Z.; Chen, D.; Sun, Y.; Wang, Y. High catalytic activity for formaldehyde oxidation of AgCo/APTES@MCM-41 prepared by two steps method. Appl. Catal. A Gen. 2014, 487, 100–109. [Google Scholar] [CrossRef]
  121. Pârvulescu, V.; Tablet, C.; Anastasescu, C.; Su, B.L. Activity and stability of bimetallic Co (V, Nb, La)-modified MCM-41 catalysts. Catal. Today 2004, 93–95, 307–313. [Google Scholar] [CrossRef]
  122. Djinović, P.; Ristić, A.; Žumbar, T.; Dasireddy, V.D.B.C.; Rangus, M.; Dražić, G.; Popova, M.; Likozar, B.; Zabukovec Logar, N.; Novak Tušar, N. Synergistic effect of CuO nanocrystals and Cu-oxo-Fe clusters on silica support in promotion of total catalytic oxidation of toluene as a model volatile organic air pollutant. Appl. Catal. B Environ. 2020, 268, 118749. [Google Scholar] [CrossRef]
  123. Abdullah, A.Z.; Bakar, M.Z.A.; Bhatia, S. Combustion of chlorinated volatile organic compounds (VOCs) using bimetallic chromium-copper supported on modified H-ZSM-5 catalyst. J. Hazard. Mater. 2006, 129, 39–49. [Google Scholar] [CrossRef] [PubMed]
  124. Shayegan, Z.; Lee, C.S.; Haghighat, F. TiO2 photocatalyst for removal of volatile organic compounds in gas phase—A review. Chem. Eng. J. 2018, 334, 2408–2439. [Google Scholar] [CrossRef]
  125. Mo, J.; Zhang, Y.; Xu, Q.; Lamson, J.J.; Zhao, R. Photocatalytic purification of volatile organic compounds in indoor air: A literature review. Atmos. Environ. 2009, 43, 2229–2246. [Google Scholar] [CrossRef]
  126. Tsang, C.H.A.; Li, K.; Zeng, Y.; Zhao, W.; Zhang, T.; Zhan, Y.; Xie, R.; Leung, D.Y.C.; Huang, H. Titanium oxide based photocatalytic materials development and their role of in the air pollutants degradation: Overview and forecast. Environ. Int. 2019, 125, 200–228. [Google Scholar] [CrossRef] [PubMed]
  127. Parrino, F.; Loddo, V.; Augugliaro, V.; Camera-Roda, G.; Palmisano, G.; Palmisano, L.; Yurdakal, S. Heterogeneous photocatalysis: Guidelines on experimental setup, catalyst characterization, interpretation, and assessment of reactivity. Catal. Rev. 2019, 61, 163–213. [Google Scholar] [CrossRef]
  128. Parrino, F.; Bellardita, M.; García-López, E.I.; Marcì, G.; Loddo, V.; Palmisano, L. Heterogeneous photocatalysis for selective formation of high-value-added molecules: Some chemical and engineering aspects. ACS Catal. 2018, 8, 11191–11225. [Google Scholar] [CrossRef]
  129. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental applications of semiconductor photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  130. Sciré, S.; Palmisano, L. Cerium Oxide (CeO₂): Synthesis, Properties and Applications; Elsevier: Amsterdam, The Netherlands, 2020. [Google Scholar]
  131. Fiorenza, R.; Bellardita, M.; Palmisano, L.; Scirè, S. A comparison between photocatalytic and catalytic oxidation of 2-Propanol over Au/TiO2-CeO2 catalysts. J. Mol. Catal. A Chem. 2016, 415, 56–64. [Google Scholar] [CrossRef]
  132. Fiorenza, R.; Condorelli, M.; D’Urso, L.; Compagnini, G.; Bellardita, M.; Palmisano, L.; Scirè, S. Catalytic and photothermo-catalytic applications of TiO2-CoOx composites. J. Photocatal. 2020, 1, 1. [Google Scholar] [CrossRef]
  133. Bellardita, M.; Fiorenza, R.; Palmisano, L.; Scirè, S. Photocatalytic and photothermocatalytic applications of cerium oxide-based materials. In Cerium Oxide (CeO₂): Synthesis, Properties and Applications; Elsevier: Amsterdam, The Netherlands, 2020; pp. 109–167. [Google Scholar]
  134. Higashimoto, S.; Tanihata, W.; Nakagawa, Y.; Azuma, M.; Ohue, H.; Sakata, Y. Effective photocatalytic decomposition of VOC under visible-light irradiation on N-doped TiO2 modified by vanadium species. Appl. Catal. A Gen. 2008, 340, 98–104. [Google Scholar] [CrossRef]
  135. Fiorenza, R.; Bellardita, M.; Scirè, S.; Palmisano, L. Effect of the addition of different doping agents on visible light activity of porous TiO2 photocatalysts. Mol. Catal. 2018, 455, 108–120. [Google Scholar] [CrossRef]
  136. Fiorenza, R.; Di Mauro, A.; Cantarella, M.; Gulino, A.; Spitaleri, L.; Privitera, V.; Impellizzeri, G. Molecularly imprinted N-doped TiO2 photocatalysts for the selective degradation of o-phenylphenol fungicide from water. Mater. Sci. Semicond. Process. 2020, 112, 105019. [Google Scholar] [CrossRef]
  137. Sytwu, K.; Vadai, M.; Dionne, J.A. Bimetallic nanostructures: Combining plasmonic and catalytic metals for photocatalysis. Adv. Phys. X 2019, 4, 1619480. [Google Scholar] [CrossRef] [Green Version]
  138. Arifin, K.; Majlan, E.H.; Wan Daud, W.R.; Kassim, M.B. Bimetallic complexes in artificial photosynthesis for hydrogen production: A review. Int. J. Hydrogen Energy 2012, 37, 3066–3087. [Google Scholar] [CrossRef]
  139. Verbruggen, S.W. TiO2 photocatalysis for the degradation of pollutants in gas phase: From morphological design to plasmonic enhancement. J. Photochem. Photobiol. C Photochem. Rev. 2015, 24, 64–82. [Google Scholar] [CrossRef]
  140. Fiorenza, R.; Bellardita, M.; D’Urso, L.; Compagnini, G.; Palmisano, L.; Scirè, S. Au/TiO2-CeO2 catalysts for photocatalytic water splitting and VOCs oxidation reactions. Catalysts 2016, 6, 121. [Google Scholar] [CrossRef] [Green Version]
  141. Colmenares, J.C.; Lisowski, P.; Łomot, D.; Chernyayeva, O.; Lisovytskiy, D. Sonophotodeposition of bimetallic photocatalysts Pd-Au/TiO2: Application to selective oxidation of methanol to methyl formate. ChemSusChem 2015, 8, 1676–1685. [Google Scholar] [CrossRef]
  142. Czelej, K.; Cwieka, K.; Colmenares, J.C.; Kurzydlowski, K.J.; Xu, Y.J. Toward a comprehensive understanding of enhanced photocatalytic activity of the bimetallic PdAu/TiO2 catalyst for selective oxidation of methanol to methyl formate. ACS Appl. Mater. Interfaces 2017, 9, 31825–31833. [Google Scholar] [CrossRef]
  143. Cybula, A.; Nowaczyk, G.; Jarek, M.; Zaleska, A. Preparation and characterization of Au/Pd Modified-TiO2 photocatalysts for phenol and toluene degradation under visible light—The effect of calcination temperature. J. Nanomater. 2014, 2014, 918607. [Google Scholar] [CrossRef]
  144. Zhang, Y.; Park, S.J. Au–pd bimetallic alloy nanoparticle-decorated BiPO4 nanorods for enhanced photocatalytic oxidation of trichloroethylene. J. Catal. 2017, 355, 1–10. [Google Scholar] [CrossRef]
  145. Zhang, Y.; Park, S.J. Bimetallic AuPd alloy nanoparticles deposited on MoO3 nanowires for enhanced visible-light driven trichloroethylene degradation. J. Catal. 2018, 361, 238–247. [Google Scholar] [CrossRef]
  146. Han, C.; Yang, X.; Gao, G.; Wang, J.; Lu, H.; Liu, J.; Tong, M.; Liang, X. Selective oxidation of methanol to methyl formate on catalysts of Au-Ag alloy nanoparticles supported on titania under UV irradiation. Green Chem. 2014, 16, 3603–3615. [Google Scholar] [CrossRef]
  147. Zielińska-Jurek, A.; Zaleska, A. Ag/Pt-modified TiO2 nanoparticles for toluene photooxidation in the gas phase. Catal. Today 2014, 230, 104–111. [Google Scholar] [CrossRef]
  148. Wysocka, I.; Markowska-Szczupak, A.; Szweda, P.; Ryl, J.; Endo-Kimura, M.; Kowalska, E.; Nowaczyk, G.; Zielińska-Jurek, A. Gas-phase removal of indoor volatile organic compounds and airborne microorganisms over mono- and bimetal-modified (Pt, Cu, Ag) titanium(IV) oxide nanocomposites. Indoor Air 2019, 29, 979–992. [Google Scholar] [CrossRef] [PubMed]
  149. Wolski, L.; El-Roz, M.; Daturi, M.; Nowaczyk, G.; Ziolek, M. Insight into methanol photooxidation over mono- (Au, Cu) and bimetallic (AuCu) catalysts supported on niobium pentoxide—An operando-IR study. Appl. Catal. B Environ. 2019, 258, 117978. [Google Scholar] [CrossRef]
Figure 1. Possible morphologies of bimetallic nanoparticles (one metal in red, the other in dark yellow): (a,b) core–shell system; (c) multishell system; (d) subcluster segregated systems; and ordered (e) and random (f) homogeneous alloys. Reprinted (adapted) from [58], Copyright 2008, American Chemical Society.
Figure 1. Possible morphologies of bimetallic nanoparticles (one metal in red, the other in dark yellow): (a,b) core–shell system; (c) multishell system; (d) subcluster segregated systems; and ordered (e) and random (f) homogeneous alloys. Reprinted (adapted) from [58], Copyright 2008, American Chemical Society.
Catalysts 10 00661 g001
Figure 2. Lattice ceria oxygens of Au/CeO2 catalyst involved in catalytic oxidation of formaldehyde. Reprinted (adapted) from [80], Copyright 2014, American Chemical Society.
Figure 2. Lattice ceria oxygens of Au/CeO2 catalyst involved in catalytic oxidation of formaldehyde. Reprinted (adapted) from [80], Copyright 2014, American Chemical Society.
Catalysts 10 00661 g002
Figure 3. Catalytic oxidation of toluene under ageing condition over Au-Pd/Nb-TiO2 catalyst. Figure from [86].
Figure 3. Catalytic oxidation of toluene under ageing condition over Au-Pd/Nb-TiO2 catalyst. Figure from [86].
Catalysts 10 00661 g003
Figure 4. TEM (Transmission electron microscope) and HRTEM (High-resolution transmission electron microscope), images of: Au-Pd/CeO2-rod (a,b); Au-Pd/CeO2-polyhedron (c,d); and Au-Pd/CeO2-cube (e,f) catalysts. Figure from Ref. [90], Copyright 2020, Elsevier.
Figure 4. TEM (Transmission electron microscope) and HRTEM (High-resolution transmission electron microscope), images of: Au-Pd/CeO2-rod (a,b); Au-Pd/CeO2-polyhedron (c,d); and Au-Pd/CeO2-cube (e,f) catalysts. Figure from Ref. [90], Copyright 2020, Elsevier.
Catalysts 10 00661 g004
Figure 5. Temperature at which the 50% of conversion of ethanol and 2-propanol was achieved (T50) versus TPR (temperature-programmed reduction) initial temperature: (filled diamond) Au-Ag/CeO2; (filled circle) Au-Cu/CeO2; (filled square) Au/CeO2; (filled triangle) Ag/CeO2; (times) Cu/CeO2. Figure from [68], Copyright 2015, Springer Nature.
Figure 5. Temperature at which the 50% of conversion of ethanol and 2-propanol was achieved (T50) versus TPR (temperature-programmed reduction) initial temperature: (filled diamond) Au-Ag/CeO2; (filled circle) Au-Cu/CeO2; (filled square) Au/CeO2; (filled triangle) Ag/CeO2; (times) Cu/CeO2. Figure from [68], Copyright 2015, Springer Nature.
Catalysts 10 00661 g005
Figure 6. DRIFT (Diffuse Reflectance Infrared Fourier Transform Spectroscopy)-CO spectra of the mono- and bimetallic Ru-Au/TiO2 samples at room temperature. Figure from [100]. Copyright 2017, Elsevier.
Figure 6. DRIFT (Diffuse Reflectance Infrared Fourier Transform Spectroscopy)-CO spectra of the mono- and bimetallic Ru-Au/TiO2 samples at room temperature. Figure from [100]. Copyright 2017, Elsevier.
Catalysts 10 00661 g006
Figure 7. (a) Ethanol oxidation mechanism on 0.5 wt% Pt/MCM-41; and (b) ethanol oxidation mechanism on 0.5 wt% Pt–15 wt% Co/MCM-41. Figure modified from [114]. Copyright 2012, Elsevier.
Figure 7. (a) Ethanol oxidation mechanism on 0.5 wt% Pt/MCM-41; and (b) ethanol oxidation mechanism on 0.5 wt% Pt–15 wt% Co/MCM-41. Figure modified from [114]. Copyright 2012, Elsevier.
Catalysts 10 00661 g007
Figure 8. Formaldehyde oxidation on monometallic Ag/MCM-41, Co/MCM-41 and bimetallic Ag-Co/MCM-41. Figure from [120]. Copyright 2014, Elsevier.
Figure 8. Formaldehyde oxidation on monometallic Ag/MCM-41, Co/MCM-41 and bimetallic Ag-Co/MCM-41. Figure from [120]. Copyright 2014, Elsevier.
Catalysts 10 00661 g008
Figure 9. Sonophotodeposition setup before (a) and during the deposition procedure (b): (1) batch photoreactor; (2) argon line; (3) switched off 6 W UV lamp; (4) ultrasonic bath; (5) switched on 6W UV lamp; (6) reflux condenser; and (7) lamp cooling system (20 °C). Figure from [141], Copyright 2015, John Wiley and Sons.
Figure 9. Sonophotodeposition setup before (a) and during the deposition procedure (b): (1) batch photoreactor; (2) argon line; (3) switched off 6 W UV lamp; (4) ultrasonic bath; (5) switched on 6W UV lamp; (6) reflux condenser; and (7) lamp cooling system (20 °C). Figure from [141], Copyright 2015, John Wiley and Sons.
Catalysts 10 00661 g009
Figure 10. Reaction mechanism of the trichloroethylene oxidation with the Au-Pd/BiPO4 photocatalyst. Figure from [144]. Copyright 2017, Elsevier.
Figure 10. Reaction mechanism of the trichloroethylene oxidation with the Au-Pd/BiPO4 photocatalyst. Figure from [144]. Copyright 2017, Elsevier.
Catalysts 10 00661 g010
Table 1. Comparison between different supported Au-Pd bimetallic catalysts in catalytic oxidation of various volatile organic compounds (VOCs).
Table 1. Comparison between different supported Au-Pd bimetallic catalysts in catalytic oxidation of various volatile organic compounds (VOCs).
Catalyst 1Preparation MethodSupportVOCT90 (°C)Ref.
1%Au-0.5%Pdcore–shellTiO2toluene≈200 °C[85]
1%Au-1%Pdchemical reductionMnO2toluene≈180 °C[87]
1%Au-1%Pdchemical reductionCo3O4toluene≈160 °C[88]
1%Au-0.5%Pdcore–shellTiO2propene≈190 °C[85]
3%Au-1%Pddeposition–precipitationCeO2-5%Fe2O3benzene≈95 °C[89]
1%Au-1%Pddeposition–precipitationCeO2benzyl alcohol≈120 °C[90]
2%Au-2%Pd-0.2%Fechemical reductionMn2O3o-xylene≈210 °C[91]
1 Nominal concentration, weight percentage (wt%).
Table 2. Comparison between different supported bimetallic catalysts in catalytic oxidation of various VOCs.
Table 2. Comparison between different supported bimetallic catalysts in catalytic oxidation of various VOCs.
Catalyst 1Preparation MethodSupportVOCT90 (°C)Ref.
0.2%Pt-0.1%PdhydrothermalSilica MCM-41toluene≈170 °C[105]
0.3%Pt-2%Pdimpregnationγ-Al2O3benzene≈225 °C[106]
2%Ru-5%CoimpregnationTiO2benzene≈200 °C[109]
1%Ru-5%CeimpregnationTiO2chlorobenzene≈275 °C[110]
18%Mn-0.1%Pdimpregnationγ-Al2O3formaldehyde/methanol≈80 °C[111]
15%Mn-5%Cuimpregnationγ-Al2O3toluene≈350 °C[112]
1.3%Fe-1.75%Agionic exchangeZMS-5ethyl acetate≈250 °C[113]
1 Nominal concentration, weight percentage (wt%).

Share and Cite

MDPI and ACS Style

Fiorenza, R. Bimetallic Catalysts for Volatile Organic Compound Oxidation. Catalysts 2020, 10, 661. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10060661

AMA Style

Fiorenza R. Bimetallic Catalysts for Volatile Organic Compound Oxidation. Catalysts. 2020; 10(6):661. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10060661

Chicago/Turabian Style

Fiorenza, Roberto. 2020. "Bimetallic Catalysts for Volatile Organic Compound Oxidation" Catalysts 10, no. 6: 661. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10060661

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