Next Article in Journal
Biotechnological Approach for the Production of Enantiomeric Hydroxylactones Derived from Benzaldehyde and Evaluation of Their Cytotoxic Activity
Next Article in Special Issue
Hydrogen Production via Pd-TiO2 Photocatalytic Water Splitting under Near-UV and Visible Light: Analysis of the Reaction Mechanism
Previous Article in Journal
Pt-Co3O4 Superstructures by One-Pot Reduction/Precipitation in Bicontinuous Microemulsion for Electrocatalytic Oxygen Evolution Reaction
Previous Article in Special Issue
Application of Heterogeneous Catalysts for Biodiesel Production from Microalgal Oil—A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Origin of Au/Ce1-xZrxO2 Catalyst’s Active Sites in Low-Temperature CO Oxidation

by
Izabela Dobrosz-Gómez
1,*,
Miguel-Ángel Gómez-García
2 and
Jacek Michał Rynkowski
3
1
Grupo de Investigación en Procesos Reactivos Intensificados y Materiales Avanzados—PRISMA, Departamento de Física y Química, Facultad de Ciencias Exactas y Naturales, Universidad Nacional de Colombia, Sede Manizales, 170004 Manizales, Colombia
2
Grupo de Investigación en Procesos Reactivos Intensificados y Materiales Avanzados—PRISMA, Departamento de Ingeniería Química, Facultad de Ingeniería y Arquitectura, Universidad Nacional de Colombia, Sede Manizales, 170004 Manizales, Colombia
3
Institute of General and Ecological Chemistry, Lodz University of Technology, Żeromskiego 116, 90-924 Lodz, Poland
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(11), 1312; https://doi.org/10.3390/catal10111312
Submission received: 1 October 2020 / Revised: 28 October 2020 / Accepted: 9 November 2020 / Published: 13 November 2020
(This article belongs to the Special Issue Commemorative Issue in Honor of Professor Hugo de Lasa)

Abstract

:
Gold catalysts have found applications in many reactions of both industrial and environmental importance. Great interest has been paid to the development of new processes that reduce energy consumption and minimize pollution. Among these reactions, the catalytic oxidation of carbon monoxide (CO) is an important one, considering that a high concentration of CO in the atmosphere creates serious health and environmental problems. This paper examines the most important achievements and conclusions arising from the own authorship contributions concerning (2 wt. % Au)/Ce1−xZrxO2 catalyst’s active sites in low-temperature CO oxidation. The main findings of the present review are: (1) The effect of preparing conditions on Au crystallite size, highlighting some of the fundamental underpinnings of gold catalysis: the Au surface composition and the poisoning effect of residual chloride on the catalytic activity of (2 wt. % Au)/Ce1−xZrxO2 catalysts in CO oxidation; (2) The identification of ion clusters related to gold and their effect on catalyst’ surface composition; (3) The importance of physicochemical properties of oxide support (e.g., its particle size, oxygen mobility at low temperature and redox properties) in the creation of catalytic performance of Au catalysts; (4) The importance of oxygen vacancies, on the support surface, as the centers for oxygen molecule activation in CO reaction; (5) The role of moisture (200–1000 ppm) in the generation of enhanced CO conversion; (6) The Au-assisted Mars-van Krevelen (MvK) adsorption–reaction model was pertinent to describe CO oxidation mechanism. The principal role of Au in CO oxidation over (2 wt. % Au)/Ce1−xZrxO2 catalysts was related to the promotion in the transformation process of reversibly adsorbed or inactive surface oxygen into irreversibly adsorbed active species; (7) Combination of metallic gold (Au0) and Au-OH species was proposed as active sites for CO adsorption. These findings can help in the optimization of Au-containing catalysts.

Graphical Abstract

1. Introduction

It is a great pleasure and honor for us to contribute to this Special Issue of the Catalysts journal dedicated to recognizing the outstanding career contributions of Professor Hugo de Lasa to the fields of heterogeneous catalysis, photocatalysis and catalytic reaction engineering. Over the years, in the subject of heterogeneous catalysis, Professor de Lasa has defined a clear path for the application of several types of catalysts, including, but not limited to, nickel-based ones, HZSM-5 zeolites, TiO2 (photocatalysis) and Al2O3 to different reactive processes. Professor de Lasa was and continues to be a constant source of inspiration for both the academic and industry sectors related to the application of chemical reaction engineering.
Gold catalysts have found applications in many reactions of both industrial and environmental importance. Great interest has been paid to the development of new processes that reduce energy consumption and minimize pollution. Among these reactions, carbon monoxide (CO) oxidation is an important one [1,2]. Moreover, CO oxidation serves as a prototypical reaction for heterogeneous catalysis. CO has no color, no smell, no taste, and it is a toxic gas at high concentrations. It is a byproduct of partial combustion of fuels in conditions of oxygen deficiency, and it is one of the main pollutants originating from internal combustion engines. It is estimated that more than 66% of the synthetic CO gas present in the air originates from automotive emissions, even if current three-way catalysts (TWCs), based on noble metals, remove 99% of CO [3]. Considering its negative impact on the environment and its high toxicity to humans and animals (due to the strong affinity to hemoglobin in the bloodstream resulting in decreasing the oxygen transport capacity of blood), CO is one of the four major air pollutants regulated both in the European Union (“Euro emissions standards”) and in many other countries in the world. CO is a poison for platinum-based catalysts for proton exchange membrane fuel cells (PEMFC). Thus, before entering PEMFC, its content must be reduced to levels below 50 ppm [4]. There are several techniques suitable for CO removal, including, but not limited to, adsorption, CO methanation and catalytic oxidation. However, CO oxidation (CO + 0.5 O2 ↔CO2; ΔH˚ = −283 kJ/mol) is the most efficient one.
Au catalysts can be used for CO removal in CO2 lasers and CO gas sensors, air-purification devices for respiratory protection, pollution control devices for reducing fixed (industrial) and mobile emission or even for removing trace quantities of CO from ambient air in enclosed atmospheres [5]. The interest in studying supported gold catalysts has increased since Haruta and coworkers discovered their extraordinary activity for low-temperature CO oxidation [6]. They reported that, compared to a highly dispersed supported Pt catalyst, the Au catalyst is much more active at temperatures below 130 °C [7]. Attention has been raised not only from the commercial implication of this discovery but also from the desire to answer the questions of why normally inert Au material can catalyze chemical reactions so efficiently.
Most studies on Au catalysts have suggested that their activity is highly sensitive to the details of the preparation procedure and catalytic testing conditions [8,9,10,11,12]. Metal-support interactions have a significant effect on the performance of supported gold catalysts. Additionally, their efficiency in CO oxidation at low temperatures depends on various factors, including the nature of the support, Au particle size, preparation procedure, pretreatment conditions and properties of the gold-support interface. A majority of preparation methods derive from the deposition–precipitation technique proposed by Haruta´s team [13]. High activity can originate at the gold-support interface, with the support playing a crucial role. Stable activity can result from optimizing solution aging during the preparation and low calcination temperatures. The deposition–precipitation using urea (CO(NH2)2) may be advantageous over that using NaOH [14]. Catalysts prepared using conventional impregnation result to be less active than the platinum group metal catalysts prepared in this way [15], especially when chloroauric acid is used as a precursor. The grafting methods use monodispersed Au colloids stabilized by organic ligands or polymer compounds [16]. Only the gas phase grafting technique is effective for the deposition of Au nanoparticles on activated carbon, but the size of Au particles is relatively large. Direct anionic exchange (DAE) provides a small gold particle size on Al2O3 [17] and hydrotalcite [18], both with high catalytic activity in CO oxidation.
CeO2 is one of the most studied active supports due to its redox behavior achieved by easy and reversible transformation of Ce4+ cations into Ce3+ ones, accompanied by oxygen vacancies formation. There are several reasons for CeO2 application as a catalyst component: (i) a noble metal itself, used as catalyst, is economically unsustainable (e.g., its high price); (ii) CeO2 exhibits high oxygen storage (release) capacity (OSC) due to its redox properties, comparing with various metal oxides (TiO2, Fe2O3, SnO2, Co3O4, MnOx and Al2O3); (iii) the OSC guarantees stoichiometric composition of gases during air-to-fuel oscillation by releasing and storing oxygen under fuel-lean and fuel-rich conditions, respectively; (iv) a beneficial effect of precious metal—CeO2 interactions on the activity of the redox couple Ce4+/Ce3+ with its ability to change from Ce4+ under oxidizing conditions to Ce3+ under net reducing conditions and vice versa; (v) CeO2 is known to affect the dispersion of supported metals, promote noble metal reduction and oxidation, store and release oxygen, hydrogen and sulfur, form surface and bulk vacancies; (vi) CeO2 presents oxygen vacancies giving it the ability to exchange oxygen species with the environment, converting it into a kind of an oxygen buffer with high oxygen storage capacity [19,20].
However, pure CeO2 does not catalyze CO oxidation at low temperatures, and it is characterized by poor thermal stability [21,22]. It can be improved by (i) doping of CeO2 fluorite lattice with metals that enhance the intrinsic activity and/or physical properties (e.g., the presence of foreign cations such as Al3+, Si4+ or Zr4+ in the CeO2 lattice, forming a solid solution, significantly improves the stability of the surface area and strongly enhance the redox properties of ceria) and (ii) loading of metal nanoparticles on CeO2 surface that provides novel active interfacial sites between metals and CeO2 surface. Thus, the expected synergetic effect between (noble) metallic nanoparticles and CeO2-based materials is a promising methodology to develop active and stable catalysts for low-temperature CO oxidation [23].
Since our earlies work [21,24], we have demonstrated the effectiveness of CeO2–ZrO2 mixed oxides used as supports for Au nanoparticles in low-temperature CO oxidation. Their unique redox properties and high mobility of lattice oxygen, strongly depending on the Ce/Zr molar ratio, were important arguments to select CeO2–ZrO2 mixed oxides as support for Au nanoparticles. Furthermore, CeO2 has already been known to be active for the removal of post-combustion pollutants [20]. Moreover, the last generation of commercial TWCs presented better catalytic performance when the supports contain ceria-zirconia solid solutions instead of pure CeO2, as reported by Daturi et al. [25]. Finally, a similar value of the isoelectric point (IEP) for CeO2, ZrO2 and consequently, their mixed oxides allowed to create comparable conditions for gold deposition.
During the past decade, the latest research trends on CO oxidation are focused on several topics. The first one is related to the reduction of the high cost of noble metals and to the improvement of catalyst stability at high temperatures [26,27,28]. Thus, the application of single-atom catalysts (SACs), e.g., Au/CeO2, able to convert CO into CO2 at low temperatures, and characterized with good stability and a low price, has recently drawn considerable attention [29]. SACs unique structure can contribute to activating the adsorbed CO and O2 and stabilizing the intermediates. Electron transfer between the metal atom and the support can play an important role in tuning the electronic structure, which can influence their activity, selectivity and stability [30]. The second investigation trend is related to the importance of understanding the dynamic behavior of catalytic oxidation of CO for the future development of catalytic systems [31]. Finally, the last group of topics is related to unveil the mechanism of catalytic CO oxidation. Although this reaction is considered as a relatively simple one, over 20 different mechanistic steps are proposed in the open literature [32,33,34,35,36,37,38]. In general, they are based on two widely accepted possible reaction pathways: Langmuir–Hinshelwood (LH) and Mars van Krevelen (MvK) [39,40,41]. The LH dual-site mechanism concerns competitive adsorption of CO and oxygen molecules from the gas phase on the catalyst surface, their subsequent reaction and CO2 desorption to the gas phase [42]. On the other hand, in the MvK mechanism, surface lattice oxygen close to the Au nanoparticles represents the active oxygen species for CO oxidation, which reacts with CO adsorbed on the Au nanoparticles. Consequently, the reaction proceeds via the continuous formation and replenishment of surface oxygen vacancies by the reaction of the catalyst with CO and O2 (Au-assisted MvK mechanism) [43].
Here, as an extension of our previous works [21,22,23,24,44,45,46,47,48], the detailed insight into the origin of Au/Ce1−xZrxO2 catalyst’s active sites in low-temperature CO oxidation is reviewed. Special attention is paid to the factors that influence their high activity, e.g., Au crystallite size, presence of residual chloride, catalyst surface composition, support composition, catalysts pretreatment, the relation between the physicochemical properties of the supports and the catalytic performance of Au/Ce1−xZrxO2. The catalyst’s reducibility is used as a tool to determine both the support and the Au particles’ chemical state and their contribution to the reaction. A conclusive analysis of the CO oxidation mechanism over Au/Ce1−xZrxO2 catalysts is also presented.

2. Factors Affecting CO Oxidation Reaction

2.1. Effect of the Preparation Conditions and the Catalyst Properties

Details of DAE preparation procedure, see Materials and Methods section, related to the application of the washing procedure for residual chloride removal from the catalysts, significantly affect (2 wt. % Au)/Ce1−xZrxO2 catalysts properties. Residual chloride presents a poisonous effect on Au supported on Al2O3 or TiO2 [49]. It can affect the catalytic activity in two different ways: (i) facilitating the agglomeration of Au particles during the calcination step by the formation of Au-Cl-Au bridges and increasing the average gold particle size; and (ii) inhibiting the catalytic activity by poisoning the active site. Its quantitative effect depends on the sample composition. The amount of adsorbed and/or incorporated chloride depends on the nature of the support, its isoelectric point and its adsorption capacity. The high affinity of Cl for the Au active site suggests that the active site is partly cationic in nature. This section details the effect of the applied washing procedure on the properties of (2 wt. % Au)/Ce1−xZrxO2 catalysts and their performance in CO oxidation. Specifically, the real Au content, the Au particle size, the presence of residual chloride, catalyst crystallinity and morphology, and catalyst surface composition are reviewed [21,24].
Figure 1 shows that the washing procedure applied to remove residual chlorine during the catalyst preparation stage affects both the real Au content and Au particle size.
All catalysts prepared without washing show the real Au content almost identical to the nominal one (2 wt. %), confirming high efficiency of the DAE preparation procedure. Only, for Au/ZrO2, the real Au-loading is ca. 20% lower than the expected one. Considering strong Lewis-type acidity of ZrO2, which is beneficial to an anionic exchange, we believe that this is related to the significantly lower specific surface area of ZrO2 (4.5 m2/g), comparing to CeO2 (58.1 m2/g) and Ce1−xZrxO2 oxides (50.1 m2/g, 46.5 m2/g and 41 m2/g for x equaled to 0.25, 0.5 and 0.75, respectively). Notice that subsequent washing of the catalyst with warm water or ammonia led to the loss of Au (ca. 5–10% for the catalysts washed with warm water and ca. 10–15% for catalysts washed with ammonia). This is due to: (i) the removal of non-attached gold complexes, simply adsorbed on the support surface due to the deficient quantity of hydroxyl groups available for anionic exchange; and (ii) the replacement of Cl ligands by OH groups in the exchanged Au species [50]. We have already reported on the high efficiency of ammonia for residual chloride ions removal (e.g., their content in Au/Mg4Al2 [18] was found to be lower than 200 ppm). Independent of the washing procedure, the specific surface area of (2 wt. % Au)/Ce1−xZrxO2 catalyst is similar to that of the corresponding oxide support (Figure 1) [24].
The effect of the washing procedure on Au particle size is also resumed in Figure 1. Catalysts washed with either warm water or ammonia present Au particle size ca. 30 or 70% lower than that of catalysts prepared without washing. BSE photomicrographs, EDS maps and characteristic X-ray spectra of (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts, prepared using different washing procedures, are presented in Figure 2. Notice that the uncalcined catalyst, prepared without washing (Figure 2a), presents a more homogeneous distribution of Au than the calcined one (Figure 2b). This last one shows characteristic microareas with higher Au concentration, resulting from Au–Cl–Au bridges formation during the calcination process [51]. A similar observation was made basing on HRTEM results (Figure 3a). For Au/CeO2 and Au/Ce0.75Zr0.25O2 catalysts, prepared without washing, most Au particles ranged from 5 to 30 nm. For Au/ZrO2, they were in the range of 10–35 nm. However, washing the catalysts with warm water leads to a more homogeneous distribution of Au (Figure 2c) and a decrease in the Au particle size (Figure 3b). Thus, for Au/CeO2 and Au/Ce0.75Zr0.25O2 catalysts, Au particles were in the range of 5–15 nm. For Au/ZrO2, they ranged from 5 to 20 nm. Still, Au agglomerates larger than 20 nm were observed. This due to the incomplete removal of residual chlorine from the catalysts, using warm water as a washing agent. On the other hand, for the catalyst washed with ammonia (Figure 2d), neither Au agglomerations nor Au clusters were found. Moreover, the signal characteristic of Cl, observed in the X-ray spectra of catalysts prepared without washing (Figure 2a,b), is absent in the spectra of washed catalysts (Figure 2c,d). For ammonia washed Au/CeO2, and Au/Ce0.75Zr0.25O2, mostly spherical Au particles were observed in the range of 1–5 nm. For Au/ZrO2, Au particle size ranged from 1 to 7 nm (Figure 3c). Thus, Au particle agglomeration, due to the presence of residual chlorine originated from HAuCl4.3H2O, can be successfully avoided using ammonia as a washing agent.
The ammonia washing procedure was more efficient for residual chlorine removal for (2 wt. % Au)/ZrO2 than for both (2 wt. % Au)/CeO2 and (2 wt. % Au)/Ce0.75Zr0.25O2 (Figure 4). Moreover, for CeO2-containing catalysts, both prepared without washing and washed with ammonia, the presence of Cl—containing species was more evident. This is due to a different mechanism of chloride ions incorporation into these catalysts. The Cl can be present both in the exchanged Au species and in the support lattice. According to Force et al. [49], chloride ions can be incorporated into an oxygen vacancy of ceria (Lewis-type sites) and/or it can be exchanged with ceria’s OH groups (Brønsted-type sites). For this reason, for catalyst supported on ZrO2, lower content of Cl—containing species were observed.
We used the ToF-SIMS technique to determine the effect of the washing procedure on the catalyst’s surface composition [24]. Figure 4 summarizes the values of ToF-SIMS selected secondary ions intensity ratios for some (2 wt. % Au)/Ce1−xZrxO2 catalysts prepared both without washing and washed with ammonia. Ion clusters related to gold were found. For all catalysts, the presence of Cl ions in the surrounding of Au centers (e.g., AuCl, AuCl2, etc.) was confirmed. Catalysts prepared without washing presented the intensity ratio of Cl ions to the total number of ions (Cl/total) significantly higher than that for the catalysts washed with ammonia. Since the average Au particle size of catalysts prepared without washing was also significantly higher than that of systems washed with ammonia (Figure 3), the role of residual chlorine in the sintering of Au can be accepted. Moreover, for Ce-containing catalysts prepared without washing, the intensity ratio of (Cl/total), (AuCl/Au), (AuCl2/Au) and (AuClO/Au) was significantly higher than that for (2 wt. % Au)/ZrO2. This observation confirms a higher affinity of ceria to chlorine.
For all studied catalysts, the presence of oxygen ions in the surrounding of gold centers (AuO, AuO2, AuOH, etc.) was also confirmed [21,24,52]. Moreover, the intensity ratio of (AuO/Au), (AuOH/Au) and (AuO2/Au) was higher for Au supported on Ce-containing oxides. Fu et al. [53], using the XPS method, confirmed the presence of different Au-containing ion clusters on Au/γ-Al2O3, Au/TiO2 and on a gold foil. However, they detected only metallic Au and AuCl ion fragments. Any ion fragments related to AuO, AuOH and AuO2 were found. Thus, the presence of AuO, AuO2 and AuOH ion clusters points to the existence of different oxidized species of gold on (2 wt. % Au)/Ce1−xZrxO2 surface. This suggests either partial oxidation of Au or chemisorption of oxygen on gold nanoparticles. Moreover, the presence of AuOH ion clusters indicates the chemisorption of water on Au particles [53].
The morphology of (2 wt. % Au)/Ce1−xZrxO2 catalysts was highly uniform. It strongly depended on Ce1−xZrxO2 composition [21], as confirmed by SEM–EDS technique (Figure 5). The interaction of Au precursor (HAuCl4.3H2O) with CeO2 during the catalyst preparation does not lead to the aggregation of spherical grains of the oxide support (Figure 5a). Neither agglomeration of regular and plated shaped particles of Ce0.75Zr0.25O2 (Figure 5b) nor angular particles of ZrO2 was observed (Figure 5c). For all catalysts, Au was homogenously distributed on the supports surface. In general, the distribution of all elements was very uniform. Neither Au clusters nor the highly concentrated Au microareas were observed.
Figure 6a displays X-ray patterns of CeO2 and (2 wt. % Au)/CeO2 catalysts, prepared without washing and washed with ammonia. Both support and catalysts exhibited four well-pronounced diffraction peaks at 2θ = 28.55°, 33.09°, 47.49° and 56.35°. They correspond to (101) (110) (200) (211) reflections, typical of face-centered cubic (fcc) cell (JCPDS: 03-065-2975) fluorite-type structure. XRD pattern of the catalyst prepared without washing also presents clear Au (111), Au (200) and Au (220) reflections at 2θ = 38.19°, 44.39°, 64.58° (JCPDS: 00-004-0784). The average Au particle size, calculated based on the Scherrer equation using (111) diffraction peak, was 15.0 nm. This result is in agreement with the result of HRTEM (dm = 15.9 nm) (Figure 3a). For (2 wt. % Au)/CeO2 catalyst prepared without washing, the crystallite size of CeO2, calculated for cerianite (111) plane, equaled to 12.3 nm, similar to that of the unsupported ceria (14.2 nm). Notice that for (2 wt. % Au)/CeO2, the observed diffraction peaks of ceria are shifted towards lower 2θ values (Figure 6a). The observed shift is even more pronounced for the catalysts washed with ammonia. We believe that in the presence of well-dispersed Au particles, during the catalyst calcination process, the autoreduction of CeO2-containing catalysts can take place. This involves the reduction of Ce4+ to Ce3+, which provokes lattice expansion due to a bigger radius of Ce4+ cation (1.14 Å vs. 0.97 Å) [20]. This is also probably the reason for the observed change in catalysts color; in the drying process from yellow to dark gray and in the calcination from dark gray to graphite color. The (2 wt. % Au)/CeO2 catalyst, washed with ammonia, presents the crystallite size of CeO2, calculated for cerianite (111) plane, of 6.4 nm, ca. twice lower than that calculated for the catalyst prepared without washing. Finally, for the catalyst washed with ammonia, Au (111), Au (220) and Au (220) reflections were not found.
XRD patterns of Ce1−xZrxO2 (x = 0, 0.25, 0.5, 0.75, 1) and of the series of (2 wt. % Au)/Ce1−xZrxO2 catalysts washed with ammonia, is presented in Figure 6b. Similar to the CeO2 system, Ce0.75Zr0.25O2 and (2 wt. % Au)/Ce0.75Zr0.25O2 presents the cubic phase (JCPDS: 00-028-0271). However, for Ce0.5Zr0.55O2, Ce0.25Zr0.75O2, (2 wt. % Au)/Ce0.5Zr0.5O2 and (2 wt. % Au)/Ce0.25Zr0.75O2, tetragonal phase was preferential (JCPDS: 00-038-1436). Simultaneously, comparing to CeO2, for Ce1−xZrxO2 (x = 0.25, 0.5, 0.75), the increasing amount of ZrO2 introduced into CeO2 lattice provokes: (i) a slight shift of the main diffraction peaks into higher 2θ values. This is a consequence of the substitution of Ce4+ cation (Ce4+ cation radius 0.97 Å) with the smaller Zr4+ one (Zr4+ cation radius 0.84 Å), causing lowering of cell symmetry. The decrease in unit cell parameters was also observed [24]; (ii) a broadening of the main diffraction peaks. This is attributed to the distortion of CeO2 cell, resulting in the decrease in the average crystallites size, as follows: for Ce0.75Zr0.25O2, for cerianite (111) plane, of 7.0 nm; for Ce0.5Zr0.55O2 and Ce0.25Zr0.75O2, for (101) plane, of 6.4 nm and 6.7 nm, respectively. For all Ce1−xZrxO2 no phase segregation was detected, confirming the successful incorporation of ZrO2 into CeO2 lattice and formation of a solid solution. No clear Au reflections were observed for (2 wt % Au)/Ce1−xZrxO2 catalysts washed with ammonia, confirming the presence of well-dispersed Au particles (ca. 4 nm, according to HRTEM, Figure 3c). Notice that, similarly to (2 wt. % Au)/CeO2 (Figure 6a), for (2 wt. % Au)/Ce1−xZrxO2 (x = 0.25, 0.5, 0.75), a characteristic shift of main diffraction peaks towards lower 2θ values (Figure 6b) is also observed. The following values of crystallite size of catalysts were determined: for (2 wt. % Au)/Ce0.75Zr0.25O2 of 7.2 nm, calculated for cerianite (111) plane, for (2 wt. % Au)/Ce0.5Zr0.5O2 and (2 wt. % Au)/Ce0.25Zr0.75O2 of 6.7 nm and 7.0 nm, respectively, calculated for (101) plane. For ZrO2, both monoclinic (JCPDS: 00-037-1484) and tetragonal phase (JCPDS: 00-050-1089) was identified. The average crystallites size of 23.4 nm, for baddeleyite (−111) plane, and of 19 nm, for (101) plane, was calculated. For (2 wt. % Au)/ZrO2, the average crystallites size of 27.8 nm and 18.8 nm was determined for (−111) and (101) plane, respectively. Moreover, any typical Au reflections were found. Thus, in accordance with HRTEM results (Figure 3c, dm = 5.6 nm), the presence of well-dispersed and small Au particles can be assumed.
It is well established that the oxidation of CO is affected by the size of the gold crystal, and, in general, the degree of CO oxidation increases with the reduction of the crystal size up to a certain size limit. Below this critical value, CO oxidation decreases (e.g., for Au/TiO2 catalyst, the reported critical value corresponds to 3.8 nm [5]). By decreasing the crystal size of Au to less than 10 nm, Haruta et coworkers found that Au oxidized CO well below 0 °C [6]. In spite of the enhancement of the catalytic activity with crystal size decreasing, this size-dependency is also representing a huge challenge in the use of catalysts with metallic nanoparticles. At the nanoscale, melting point depression occurs to the particles, increasing the mobility of metal atoms over the support surface and increasing the probability of sintering and Ostwald ripening [54]. The size of the catalyst crystals changes the contact limit between the catalyst and the support, which is considered one of the most critical factors in the oxidation of CO. The variation in the catalytic activity has been attributed to various molecular scale factors, including change of surface structure, electronic state, metal-support interactions, an active surface oxide layer (capping oxygen mobility), oxidation states, quantum confinement effects on the nanoparticle electrons, band-gap change also related to quantum effects at the nanoscale size and the increase in unsaturated surface atoms due to the nanoparticle geometry [55].
Figure 7 presents the effect of the washing procedure on the performance of (2 wt. % Au)/Ce1−xZrxO2 catalysts in CO oxidation.
For all catalysts prepared without washing, even if they present similar real gold-loading (Figure 1), significant differences in CO oxidation were observed. The (2 wt. % Au)/Ce0.5Zr0.5O2 catalyst presented the highest activity (e.g., the reaction starts at ca. 165 °C). Catalyst washing, with warm water or with ammonia, decreased Au particle size (as shown in Figure 1) and increased the catalytic activity. In comparison with catalysts prepared without washing, catalysts washed with warm water oxidized CO at lower temperatures (ca. 80 °C for the most active one). Moreover, the catalysts washed with ammonia presented a more remarkable decrease in temperature of 10% and 50% CO conversion. The highest activity was observed for (2 wt. % Au)/Ce0.75Zr0.25O2, below room temperature (e.g., 10% conversion was detected at 5 °C). Notice that for this catalyst, ToF-SIMS studies confirmed the presence of AuCl, AuClO and AuCl2 ions in the surroundings of Au particles (Figure 4), even if residual chloride concentration was lower than 200 ppm. This suggests that high CO conversion on (2 wt. % Au)/Ce1−xZrxO2 can be reached even if residual chloride ions are still present in the catalyst. Jóźwiak et al. [56] have found that Au(K)/Fe2O3 catalyst, with a Cl/Au = 4 ratio, showed considerable activity. However, the increase in the Cl/Au ratio up to 8 led to the catalyst deactivation.
Resuming, the catalytic performance of (2 wt. % Au)/Ce1−xZrxO2 depends on (i) Au particle size; (ii) Au dispersion, (ii) residual chloride content; (iv) catalyst´s surface composition and (v) support composition and properties. This last one is discussed below.

2.2. Effect of the Support Composition (Ce/Zr Molar Ratio)

The effect of support on the catalytic performance of Au catalysts starts since their synthesis. Actually, the selection of preparation procedure is related to the physicochemical properties of the support (e.g., morphology, specific surface area, acid/base properties, presence of surface defects, the possibility of phase transformation, etc.). Both for deposition–precipitation method, developed by Haruta and coworkers [6,7,13] and for the DAE technique, developed by Pitchon and coworkers [17] and used for the preparation of (2 wt. % Au)/Ce1−xZrxO2 catalysts [21,22,23,24,44,45,46,47,48], a key parameter is the surface charge of the support material. This depends on the pH value of the Au precursor and the isoelectric point (IEP) of the oxide. Their interactions proceed either by anionic or by cation adsorption. The anionic adsorption occurs when the pH of the solution is higher than the IEP of the support (the main surface species is O; thus, the surface is negatively charged). The cationic adsorption takes place when the pH is lower than the IEP of the support (the main surface species is OH2+, and the support surface is positively charged). Since the IEP of Ce1−xZrxO2 is ca. 6.7, and at pH = 2, the dominant species of Au corresponds to the negatively charged ions (AuClx(OH)4–x) [57], the anionic exchange of gold species with hydroxyl groups of the support occurs. Anion adsorption (AA) [14], cationic adsorption [58] and co-precipitation [6] methods have also been attempted by many researchers.
Supports can also affect the catalytic properties of supported metallic nanoparticles, and they can play different roles during a reaction. They can provide specific defects where Au nanoparticles can be anchored and/or stabilized. They can let the electron transfer to or from metal particles. They can provide additional functionality (acidity/basicity) to the overall supported metal catalyst. Finally, they can participate in the reaction. Herein, details on the effect of Ce1−xZrxO2 composition on catalytic properties of (2 wt. % Au)/Ce1−xZrxO2 in CO oxidation are reviewed.
In order to understand the role of support in the creation of catalytic performance of (2 wt. % Au)/Ce1−xZrxO2 in CO oxidation, TPR studies were used as a tool. Figure 8 shows TPR-H2 (a) and TPR-CO (b) profiles of Ce0.75Zr0.25O2 and (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts.
The reduction of Ce0.75Zr0.25O2 mixed oxide occurs in one stage with the maximum at 565 °C using H2 as a reducing agent. (Figure 8a). For (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts, the characteristic H2 reduction peak is shifted toward a lower temperature in the presence of Au nanoparticles characterized by different sizes. It is assigned to the co-reduction of Au3+ and of surface oxygen. The observed decrease in the reduction temperature, comparing to the unsupported mixed oxide, is related to the activation of H2 molecule on Au nanoparticles followed by the migration of dissociated hydrogen species, by a spillover process, from the surface of the Au particles to the support [44,46]. However, if the catalyst contains only ionic Au, the surface oxygen reducibility could be intensified through the lattice substitution mechanism [3,4]. According to those statements, the Au+ or Au3+ ions could fill the vacant Ce4+ sites, resulting in oxygen vacancy formations and an increase in oxygen mobility and reducibility. One can see that the smaller the Au particle size is, the bigger the shift of the reduction peak is observed. Moreover, for all (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts, a substantial decrease in TPR-H2 signal intensity is observed. Estimated H2 consumption strongly depends on Au particle size, with the minimum observed for the catalyst washed with ammonia. Considering that the washing procedures removes residual chlorine, which is also known to inhibit the reducibility of Ce-containing oxides, the catalyst autoreduction during calcination occurs [21]. Notice that amount of H2 used for the reduction of dried (uncalcined) (2 wt. % Au)/Ce0.75Zr0.25O2 catalyst, washed with ammonia, was higher (1.12 mmol/g) than that observed for the calcined catalyst, washed with ammonia (0.95 mmol/g). For comparison, the H2 consumption for Ce0.75Zr0.25O2 reduction equaled to 1.53 mmol/g.
When CO is used as a reducing agent, the reduction of Ce0.75Zr0.25O2 mixed oxide occurs in one stage with the maximum at 460 °C (Figure 8b). A characteristic shoulder is also observed at low temperature. This suggests the existence of at least two types of Cex+ located at different chemical environments, assigned to the reduction of surface and subsurface Cex+, susceptible to the reduction at a lower temperature in the presence of Au. Thus, for (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts, the characteristic CO2 formation peak consists of two overlapping peaks indicating two reduction process: (i) the first one related to the reduction of both the oxygen species coordinated around dispersed Au nanoparticles and the surface capping oxygen of support, strongly anchored with Au, and (ii) the second one, corresponding to the reduction of both oxygen species coordinated around larger Au particles and the support surface layers, not strongly anchored with Au. The smaller the Au particle size is, the bigger the shift of the reduction peak is observed. Similar to TPR-H2 profiles, the contribution of low-temperature reduction peak decreases with a decrease in Au particle size, as confirmed by HRTEM results (Figure 3), due to the catalyst autoreduction.
Figure 9 shows the correlation between the hydrogen consumption, the ratio of oxygen atoms in the capping peak to the total cerium atoms and the activity of (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts prepared using different washing procedures. The sequence of the decreasing temperature of CO oxidation indicates that the activity of Ce-containing oxides in CO oxidation depends on the amount of capping oxygen available at the lowest temperature. TPR-H2 studies showed a marked effect of well-dispersed Au nanoparticles and its size on the reducibility of Ce0.75Zr0.25O2 support at low temperatures. For Au catalysts, the H2 consumption for the removal of surface capping oxygen atoms is higher than that for the oxide support. One can see that the sequence of the increasing activity of the studied catalysts follows the increasing percentage of oxygen atoms in the capping peak to the total cerium atoms (Oc/Ce) (Figure 9).
These findings indicate the role of both Au particle size and the support redox properties in the creation of the catalytic performance of (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts. The higher reducibility of the oxide support, in the presence of well-dispersed and small Au nanoparticles, increases the concentration of surface oxygen vacancies available at low temperatures. Considering that CO adsorption was observed on both Au3+, Au+, Au0 species and CeO2 (Ce4+-CO), a synergetic effect between the support and Au nanoparticles at the interface could be implied. However, for Ce0.75Zr0.25O2 oxide, any CO conversion was observed at temperatures below 200 °C. Thus, well-dispersed Au was responsible for the improvement of its catalytic activity.
Although the nature of the active sites and the reaction mechanism is a matter of debate, there is an agreement that the activity in CO oxidation is ruled by the contact boundaries between oxide support and Au nanoparticles. Indeed, the observed differences in CO oxidation activity overstudied catalysts depend on Au particles size and their dispersion (Figure 10), since the smallest nanoparticles enhanced the Au-oxide contact boundaries formation. Notice that (2 wt. % Au)/ZrO2 catalyst presented the lowest activity as well as the biggest particle size of Au and oxide support (Figure 1 and Figure 6b). Consequently, the support particle size and reducibility (which depends on the composition) appear to be crucial in the creation of catalytic performance of supported Au catalysts in CO oxidation.
For unsupported Ce1−xZrxO2, the sequence of decreasing temperature of the maximum reduction peak clearly follows the sequence of increasing activity of these oxides in CO oxidation (Figure 10a). Ce-rich oxides (CeO2 and Ce0.75Zr0.25O2) were the most active ones. They were characterized by both a cubic crystalline phase and high reducibility at lower temperatures. Accordingly, the presence of the cubic phase in the Ce-Zr solid solution improves its catalytic activity in CO oxidation [47].
For (2 wt. % Au)/Ce1−xZrxO2 catalysts (Figure 10b–d), more homogenously dispersed Au nanoparticles increased the reducibility of the studied oxides, mostly at lower temperatures. Moreover, the initial temperature of the catalyst reduction, using CO as a reducing agent, corresponded to the T50. For details see [21,24]. This shows the role of support redox properties, in the presence of well-dispersed Au, in the creation of catalytic activity of (2 wt. % Au)/Ce1−xZrxO2 in CO oxidation. Notice that (2 wt. % Au)/Ce0.75Zr0.25O2 was the most active catalyst. It is reducible at the lowest temperature due to a higher concentration of surface oxygen vacancies (which can be the centers of oxygen activation). Such vacancies may be formed even in the reaction mixture CO + O2 exhibiting lower redox potential compared to the oxygen atmosphere, as reported by Grzybowska-Świerkosz [59] and Guzmán et al. [60]. It can be suggested that ceria, which is reducible at the lowest temperatures, appears to be responsible for the activity of these catalysts in CO oxidation.

2.3. Influence of the Pretreatment Atmosphere

The performance of Au/metal oxide catalysts in CO oxidation can be altered by the pretreatment atmosphere as it directly affects the surface composition of the catalytic system. For example, in the presence of water, an increase in the CO oxidation activity for Au supported on TiO2, Fe2O3, Al2O3, Mg(OH)2 was reported [61]. However, Olea et al. [62] informed that in the presence of water, the catalytic activity of Au/TiO2 decreases. Thus, due to the relevance on the practical applications, we studied the effect of pretreatment atmosphere (e.g., dry He, air, H2 and H2 + water, realized at 200 °C) on CO conversion at 30 °C on (2 wt. % Au)/Ce1−xZrxO2 catalysts washed with ammonia (Figure 11) [45].
The thermal treatment of all (2 wt. % Au)/Ce1−xZrxO2 in dry He significantly decreases their catalytic activity (ca. 50%) in relation to the calcined catalysts (Figure 11). However, the observed decrease in activity was re-established, exposing the catalysts, at room temperature, to the stream of moisture (6000 ppm)/He. We believe that the loss of hydroxyl species from the catalyst surface is the reason for the observed deactivation. A similar observation was also reported by Costello et al. [63]. The activity of the dried (as prepared) (2 wt. % Au)/Ce1−xZrxO2 catalysts was significantly higher than that of the calcined ones (Figure 11). Notice that the dried systems were not pretreated before activity testing. This implies that high CO oxidation yields require the presence of water-derived species on the (2 wt. % Au)/Ce1−xZrxO2 surface. Thus, the moisture adsorbed on the catalyst surface is responsible for the creation of catalytic performance rather than its content in the feed stream. However, their activity was unsteady in time and catalysts deactivation after 30 min of reaction took place. It was just partially re-established, exposing the catalysts, at room temperature, to the stream of moisture (6000 ppm)/He. The pretreatment in the H2 stream, at 200 °C, of (2 wt. % Au)/CeO2 and (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts significantly increased their catalytic activity in comparison with the calcined samples. Notice that the activity of the pre-reduced catalyst was steady in time. Any deactivation was observed after a few hours of reaction. This suggests that rather the reduction of Au nanoparticles, instead of support reduction, is responsible for the increase in activity observed over pre-reduced contacts [64]. Moreover, the pre-reduced systems showed better performance in CO oxidation than the unreduced ones. Thus, both metallic (Au°) and oxidized (Aux+) species are needed for high CO conversion at low temperatures [65,66]. On the other hand, the increase in the reduction temperature (H2 stream, at 290 °C), which allowed obtaining a higher amount of metallic Au°, provoked the decrease in activity [45]. This indicates that partial reduction of Au3+ to Au° favors high CO oxidation yields. In this case, it was achieved by reducing the catalysts at lower temperatures (H2 treatment at 200 °C). Finally, the catalyst pretreatment in an oxidative atmosphere (air) decreases catalytic activity. This probably due to the increase in the amount of Au3+ species. For (2 wt. % Au)/ZrO2, independent of the treatment atmosphere, no CO conversion was observed at low temperature.
The H2-TPR profiles of (2 wt. % Au)/Ce0.75Zr0.25O2 catalyst, washed with ammonia, pretreated in different atmospheres, are presented in (Figure 12). One can see that the reduction of each catalyst occurs in one stage. The observed reduction peak is assigned to the co-reduction of Au3+ and surface oxygen on Ce0.75Zr0.25O2. In the presence of Au, the reducibility of the Ce0.75Zr0.25O2 surface oxygen is improved, and it facilitates the oxygen transfer across the solid–gas interface during the reaction. The pretreatment in different atmospheres shifts the characteristic reduction peak to the lower temperature in the following sequence: H2 + H2O (168 °C) < dry air (175 °C) ≈H2 (177 °C) < He (185 °C) < calcined (190 °C) ≈dried catalyst (190 °C).
Moreover, a significant difference in the total H2 consumption needed for the catalyst reduction was observed. For the catalyst treated with dry He, the H2 consumption (0.93 mmol/g) is very similar to that of the calcined sample (0.95 mmol/g). This indicates that this treatment can only remove the surface impurities from (2 wt. % Au)/Ce0.75Zr0.25O2 catalyst surface. For the as-prepared (dried) catalyst, the highest value of H2 consumption was observed (1.12 mmol/g). This suggests that the dried catalyst contains mostly Au3+ and probably does not possess Au0. For the catalyst pretreated in air, H2 consumption equaled 0.99 mmol/g, indicating the presence of a considerable amount of Aux+ on the catalyst surface. The catalyst pretreated in H2 possesses a significantly higher amount of Au0 on the surface (H2 consumption 0.57 mmol/g was registered). Finally, the lowest amount of H2 (0.49 mmol/g) was used for the reduction of the most active catalyst after pretreatment in the H2 + H2O stream. Thus, the following observations with respect to observed activity can be made: (i) Two active forms of Au can be responsible for high activity in CO oxidation: Au0 and Au3+. The Au0 is crucial for CO adsorption. The Au3+ species, in conjunction with Au0, are essential for the activation of surface hydroxyl groups. Next, the activated OH groups can react with the adsorbed CO to form carboxylate species [8,9,53,65,66]. However, a high amount of Aux+ or Au° decreases the activity. In fact, the appropriate ratio of Aux+ and Au° is critical to obtain the highest catalytic activity in CO oxidation [67,68,69]; (ii) The adsorption and activation of O2 take place on the support oxygen vacancies. During CO oxidation at oxygen lean conditions, Ce-containing catalysts worked as oxygen buffers by release–uptake of oxygen [21]. (iii) In the presence of moisture, the O2 molecules can dissociate easier and can rapidly react with adsorbed CO [45]. Moreover, Au electronic state can be modified; (iv) High CO oxidation yields, on (2 wt. % Au)/Ce1−xZrxO2, require the presence of water-derived species on the catalyst surface. Thus, the moisture adsorbed on the catalyst surface, rather than its content in the feed stream, is responsible for the creation of the catalytic performance.
The obtained results suggest that CO oxidation under stationary conditions over Ce-containing oxides occurs by the Au-assisted Mars van Krevelen (MvK) redox-type mechanism. It consists of: (i) the metal/support oxidation in separate and independent steps, (ii) adsorption of CO from the gas phase on the oxide, (iii) reaction between the adsorbed CO and catalyst surface oxygen, accompanied with the creation of oxygen vacancies, (iv) CO2 desorption from the catalyst surface and (v) fast and irreversible refilling of the resulting oxygen vacancies by oxygen from the gas phase, in a separate step (as in i). The MvK model is able to predict CO oxidation experimental data at stoichiometric and rich conditions, even if some of them present a very fast increase in conversion with the reaction temperature [48]. Consequently, the principal role of Au in CO oxidation over (2 wt % Au)/Ce1−xZrxO2 catalysts could be related to the promotion in the transformation process of reversibly adsorbed or inactive surface oxygen into irreversibly adsorbed species, active in CO oxidation. On the other hand, according to the reactivity results, Figure 7, the highest quantity of oxygen that can be involved in the reaction, must be available at the same temperatures at which CO oxidation occurs. It suggests that the reaction involves both gas-phase oxygen and surface—bulk oxygen, indicating the ability of Ce-containing oxides to supply reactive oxygen to Au active species for CO oxidation. It supports the idea of Ce acting as an oxygen buffer by releasing-up taking oxygen through the redox processes involving Ce4+/Ce3+ redox couple. Similar observations were reported elsewhere [70,71,72]. Additional experiments performed under totally anaerobic conditions (without oxygen present in the gas phase), showing that the catalyst activity depends on its ability to release first the surface and then bulk oxygen [48]. Thus, as expected, the most active catalyst is (2 wt. % Au)/Ce0.75Zr0.25O2, which presents the highest oxygen mobility (Figure 9).

3. Materials and Methods

3.1. Synthesis of Catalysts

A series of Ce1−xZrxO2 (x = 0.25, 0.5, 0.75) solid solutions, CeO2 and ZrO2 used as a support for Au was prepared by the sol–gel-like method, based on thermal decomposition of mixed propionates [73]. The starting materials, zirconium (IV) acetylacetonate (Zr(CH3COCH2COCH3)4, purity 99.9%, Avocado, Tewksbury, MA, USA) and/or cerium (III) acetyloacetonatehydrate (Ce(CH3COCH2COCH3)3 H2O, purity 99.9%, Sigma–Aldrich, St. Louis, MO, USA) and calcined at 550 °C, in details described previously [21].
(2 wt. % Au)/Ce01−xZrxO2 catalysts were prepared by the direct anionic exchange (DAE) method of Au species with hydroxyl groups of support, based on a previously established procedure [21,23]. In brief, hydrogen tetrachloroaureate (III) trihydrate (HAuCl4.3H2O, purity 99.9%, Sigma–Aldrich, St. Louis, MO, USA) was used as a gold precursor. The catalysts were prepared by the direct anionic exchange (DAE) method of gold species with hydroxyl groups of the support. The optimization of catalyst synthesis conditions was presented previously [24]. An aqueous solution of HAuCl4 at the concentration of 2.25 × 10−4 M was prepared using 900 cm3 of distilled water. The solution was heated up to 70 °C, and the support was introduced. After 1 h of thermostatting and vigorous stirring, the suspension was centrifuged. In order to remove the residual chlorine from the catalysts, two washing procedures prior to the drying process were applied. Depending on the washing procedure, the solids were suspended in either 4 M ammonia solution at 25 °C or deionized water at 50 °C and stirred for 1 h and next centrifuged again. After drying in an oven at 120 °C overnight ((2 wt. % Au)/Ce1−xZrxO2 dried or as prepared), the catalysts were calcined in air at 300 °C ((2 wt. % Au)/Ce1−xZrxO2 calcined) for 4 h.

3.2. Characterization Methods

The specific surface area was determined by BET method using a Sorptomatic 1900 apparatus (Carlo-Erba, Milan, Italy) with N2 as an adsorbate. Prior to the measurement, all samples were degassed for 4 h at 250 °C.
The morphology of the prepared catalytic systems was studied using a-4700 scanning electron microscope, SEM (HITACHI, Tokyo, Japan), equipped with an energy-dispersive spectrometer, EDS (ThermoNoran, Madison, WI, USA). Before analyzing, the samples were coated with a carbon monolayer in order to reduce the charge buildup on the surface.
The amount of Au deposited on Ce0.75Zr0.25O2 was determined using an atomic absorption spectroscope (AAS), specifically a Solaar M6 Unicam spectrophotometer. Prior to the analysis, each sample was digested by aqua regia (3 cm3 + 100 mg of the sample), using a microwave oven (MLS-1200 MEGA, Milestone, Apeldoorn, The Netherlands), and the resulting solution was diluted with 100 cm3 of deionized water. A blank test was prepared in a similar way.
High-resolution transmission electron microscopy (HRTEM) measurements were carried out using high-resolution microscope EM-002B (TOPCON, Tokyo, Japan), at an acceleration voltage 200 kV, equipped with an energy-dispersive spectrometer (EDS). To determine an average Au particle size and to define their distribution more accurately, at least 500 particles of each catalyst were chosen. The mean particle diameters were calculated, as described previously [18].
X-ray diffraction patterns were obtained at room temperature using a PANalytical X’Pert Pro MPD diffractometer (Malvern, UK), operating at 40 kV and 30 mA (Cu Ka radiation). Data were collected in the range 20–70° 2θ, with a step size of 0.01678 and a step time of 10 s. JCPDS files were used for the identification of the diffraction peaks. Approximately 300 mg of sample, which had been hand ground in an agate mortar, was packed in the sample holder. The particle size was estimated from the width of principal diffraction peaks using the Scherrer equation.
Time-of-flight secondary ion mass spectrometry (ToF–SIMS) was applied to determine the chemical state of gold in the catalysts. The measurements were performed using the ION-TOF GmbH instrument (TOF–SIMS IV, Münster, Germany) equipped with 25 kV pulsed 69Ga+ primary ion gun in the static mode. The analyzed area corresponded to a square of 500 μm × 500 μm. Before the measurements, the samples were pressed into pellets and attached to the sample holder using a double-sided tape. Moreover, a pulsed electron flood gun was used for charge compensation. Thus, the peak with m/z of 197 was assigned to Au. The peaks with m/z of 213, 214, 229, 232, 248 and 267 were ascribed to ion fragments of AuO, AuOH, AuO2, AuCl, AuClO and AuCl2, respectively.
Temperature-programmed reduction (TPR-CO) measurements were carried out using the PEAK-4 apparatus, using CO as a reducing agent. The apparatus was equipped with an infrared gas analyzer (type: ZRJ-4, Fuji Electric Systems Co., Tokyo, Japan) to analyze CO2 formation. TPR-CO experiments were performed using a CO/He (5 vol % CO, 95 vol % He) gas mixture, with a flow rate of 40 cm3/min, in the temperature range 25–850 °C, with a ramp rate 15 °C/min. Powdered samples of 100 mg were exposed to dry Ar at 250 °C for 1 h before the reduction. Based on the CO2 formation, reduction degrees of the investigated samples were calculated as described previously in detail [21].
Temperature-programmed Reduction (TPR-H2) experiments were carried out using the PEAK-4 apparatus equipped with a thermal conductivity detector (TCD). They were performed using an H2/He (5 vol % H2, 95 vol % He) gas mixture, with a flow rate of 40 cm3/min, in the temperature range 25–850 °C, with a ramp rate 15 °C/min. Powdered samples of 100 mg were exposed to dry Ar at 250 °C for 1 h before reduction.
CO oxidation reaction was carried out at atmospheric pressure in a quartz flow microreactor containing 100 mg of sample in a fixed bed, using a series of mass flow controllers with diluted gases.
Prior to the catalytic tests, the samples were heated in the gas mixture stream at 300 °C for 1 h with a ramp rate of 5 °C/min, except for the pretreatment studies. In these last case, the catalysts were pretreated at 200 °C for 60 min in different atmospheres including: air (99.9 vol % Air), dry He (99.9 vol % He), H2 (5 vol % H2, 95 vol % Ar), H2 + H2O (5.0 vol %)/H2O (0.1 vol %)/Ar (94.9 vol %). The gas mixture containing 1.7 vol % CO and 3.4 vol % O2 (He as an eluent gas) was used for CO oxidation tests (flow rate of 50 cm3/min.). The activity tests were performed either in the temperature range of 25–300 °C, with a ramp rate of 5 °C/min or at isothermal conditions (at 30 °C). They were repeated several times in order to verify reproducibility. CO and O2 concentrations were analyzed using a gas chromatograph fitted with molecular sieves 5 Å and equipped with a thermal conductivity detector. CO conversion was calculated, as described previously [21]. The catalytic performance of studied materials was assessed in terms of both T10 and T50 temperatures, defined as the temperature at which 10% and 50% conversion was obtained, respectively.
The model developed by Johnson and Mooi [74,75] is based on the qualitative relationship between the magnitude of the capping oxygen and BET surface area. For this model, the authors considered that the reduction of CeO2 corresponds to the elimination of one-fourth of the surface oxygen ions and that each ceria crystallite is the cube of n oxygen ions on a side. Therefore, each ceria crystallite contains n3 oxygen ions and ½n3 Ce ions. Since the number of capping oxygen ions in a crystallite is: Oc = n3– (n − 2)3 = 6n2 − 12n + 8 and the number of H atoms to reduce half of the Oc ions is: Hc = 2·Oc, the ratio of hydrogen atoms in the capping peak to the total cerium atoms (Hc/Ce) will be: Hc/Ce = 4(6n2 − 12n + 8)/n3.

4. Conclusions

The most important achievements and conclusions arising from the own authorship contributions concerning (2 wt. % Au)/Ce1−xZrxO2 catalyst’s active sites in low-temperature CO oxidation were reviewed. They can be summarized as follows:
  • High activity of (2 wt. % Au)/Ce1−xZrxO2 catalysts in low-temperature CO oxidation is displayed by Au nanoparticles with an optimum size of ca. 4 nm. In the presence of residual chlorine, Au agglomeration during the calcination process occurs. Washing of the catalysts with ammonia effectively prevents Au sintering. The smallest Au nanoparticles and the highest activity in CO oxidation presented catalysts washed with ammonia.
  • The electronic properties of nanosized Au can be modified by the oxide support. The structure and properties of oxide support define the nature of Au-support interactions. High oxygen mobility at low temperatures is critical for high activity. The catalytic performance of (2 wt. % Au)/Ce1−xZrxO2 contacts strongly depends on the support’s redox properties. Oxygen vacancies on the support surface are centers for oxygen molecule activation.
  • Concerning Au active species: (i) Two active forms of Au can be responsible for high activity in CO oxidation: Au0 and Au3+. The Au0 is crucial for CO adsorption. The Au3+ species, in conjunction with Au0, are essential for the activation of surface hydroxyl groups. Next, the activated OH groups can react with the adsorbed CO to form carboxylate species. However, a high amount of Aux+ or Au° decreases the activity. In fact, the appropriate ratio of Aux+ and Au° is critical to obtain the highest catalytic activity in CO oxidation; (ii) The adsorption and activation of O2 take place on the support oxygen vacancies; (iii) The moisture can both enhance the dissociation of O2 molecules adsorbed on the oxygen vacancies and modify the electronic state of gold. Next, the activated oxygen can rapidly react with CO adsorbed on the metallic Au0; (iv) High CO oxidation yields require the presence of water-derived species on the catalyst surface. Thus, the moisture adsorbed on the catalyst surface, rather than its content in the feed stream, is responsible for the creation of the catalytic performance.
  • Regarding the reaction mechanism: The Au-assisted MvK model was suggested as more relevant to describe the kinetics of CO oxidation over Ce1−xZrxO2 and (2 wt. % Au)/Ce1−xZrxO2 catalysts. Consequently, the principal role of Au in CO oxidation over (2 wt. % Au)/Ce1−xZrxO2 catalysts can be related to the promotion in the transformation process of reversibly adsorbed or inactive surface oxygen into irreversibly adsorbed species active in CO oxidation. Additionally, at oxygen-lean conditions, the activity seems to be dependent on the ability of the oxide to release first the surface and then bulk oxygen. Thus, considering that extra oxygen is available from the catalyst surface bulk, it was possible to properly fit the observed catalyst activity. Finally, the proposed fitting procedure, based on MvK expression, can be considered as successful for all conditions tested for the CO oxidation reaction.

Author Contributions

Conceptualization: I.D.-G.; methodology: I.D.-G.; software, I.D.-G. and M.-Á.G.-G.; formal analysis: I.D.-G.; investigation: I.D.-G.; resources: I.D.-G., J.M.R.; writing—original draft preparation: I.D.-G.; writing—review and editing: I.D.-G., M.-Á.G.-G. and J.M.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lykaki, M.; Stefa, S.; Carabineiro, S.A.C.; Pandis, P.K.; Stathopoulos, V.N.; Konsolakis, M. Facet-Dependent Reactivity of Fe2O3/CeO2 Nanocomposites: Effect of Ceria Morphology on CO Oxidation. Catalysts 2019, 9, 371. [Google Scholar] [CrossRef] [Green Version]
  2. Soliman, N.K. Factors Affecting CO Oxidation Reaction over Nanosized Materials: A Review. J. Mater. Res. Technol. 2019, 8, 2395–2407. [Google Scholar] [CrossRef]
  3. Bond, G.C.; Thompson, D.T. Gold-catalyzed oxidation of carbon monoxide. Cat. Rev. Sci. Eng. 1999, 41, 319. [Google Scholar] [CrossRef]
  4. Hashmi, A.S.K.; Hutchings, G.J. Gold Catalysis. Angew. Chem. Int. Ed. 2006, 45, 7896–7937. [Google Scholar] [CrossRef]
  5. Du, M.; Sun, D.; Yang, H.; Huang, J.; Jing, X.; Odoom-Wubah, T.; Wang, H.; Jia, L.; Li, Q. Influence of Au particle size on Au/TiO2 catalysts for CO oxidation. J. Phys. Chem. C 2014, 118, 19150–19157. [Google Scholar] [CrossRef]
  6. Haruta, M.; Yamada, N.; Kobayashi, T.; Iijima, S. Gold catalysts prepared of coprecipitation for low-temperature oxidation of hydrogen and carbon monoxide. J. Catal. 1989, 115, 301–309. [Google Scholar] [CrossRef]
  7. Haruta, M.; Date, M. Advances in the catalysis of Au nanoparticles. Appl. Catal. A Gen. 2001, 222, 427–437. [Google Scholar] [CrossRef]
  8. Zhou, A.; Wang, J.; Wang, H.; Li, H.; Wang, J.; Shen, M. Effect of active oxygen on the performance of Pt/CeO2 catalysts for CO oxidation. J. Rare Earths. 2018, 36, 257–264. [Google Scholar] [CrossRef]
  9. Haruta, M. Size- and Support-Dependency in the Catalysis of Gold. Catal. Today 1996, 36, 153–166. [Google Scholar] [CrossRef]
  10. Cui, Y.; Xu, L.; Chen, M.; Lv, C.; Lian, X.; Wu, C.; Yang, B.; Miao, Z.; Wang, F.; Hu, X. CO Oxidation over Metal Oxide (La2O3, Fe2O3, PrO2, Sm2O3, and MnO2) Doped CuO-Based Catalysts Supported on Mesoporous Ce0.8Zr0.2O2 with Intensified Low-Temperature Activity. Catalysts 2019, 9, 724. [Google Scholar] [CrossRef] [Green Version]
  11. Bansmann, J.; Abdel-Mageed, A.M.; Chen, S.; Fauth, C.; Häring, T.; Kucerová, G.; Wang, Y.; Behm, R.J. Chemical and Electronic Changes of the CeO2 Support during CO Oxidation on Au/CeO2 Catalysts: Time-Resolved Operando XAS at the Ce LIII Edge. Catalysts 2019, 9, 785. [Google Scholar] [CrossRef] [Green Version]
  12. Sui, C.; Xing, L.; Cai, X.; Wang, Y.; Zhou, Q.; Li, M. Co-Supported CeO2 Nanoparticles for CO Catalytic Oxidation: Effects of Different Synthesis Methods on Catalytic Performance. Catalysts 2020, 10, 243. [Google Scholar] [CrossRef] [Green Version]
  13. Kaneti, Y.V.; Tanaka, S.; Jikihara, Y.; Nakayama, T.; Bando, Y.; Haruta, M. Room Temperature Carbon Monoxide Oxidation Based on Two-dimensional Gold-loaded Mesoporous Iron Oxide Nanoflakes. Chem. Commun. 2018, 54, 8514–8517. [Google Scholar] [CrossRef] [PubMed]
  14. Zanella, R.; Giorgio, S.; Henry, C.R.; Louis, C. Alternative methods for the preparation of gold nanoparticles supported on TiO2. J. Phys. Chem. B 2002, 106, 7634–7642. [Google Scholar] [CrossRef]
  15. Bamwenda, G.R.; Tsubota, S.; Nakamura, T.; Haruta, M. The influence of the preparation methods on the catalytic activity of platinum and gold supported on TiO2 for CO oxidation. Catal. Lett. 1997, 44, 83–87. [Google Scholar] [CrossRef]
  16. Grunwaldt, J.D.; Kiener, C.; Wögerbauer, C.; Baiker, A. Preparation of supported gold catalysts for low-temperature CO oxidation via “size-controlled” gold colloids. J. Catal. 1999, 181, 223–232. [Google Scholar] [CrossRef]
  17. Ivanova, S.; Petit, C.; Pitchon, V. A new preparation method for the formation of gold nanoparticles on an oxide support. Appl. Catal. A Gen. 2004, 267, 191–201. [Google Scholar] [CrossRef]
  18. Dobrosz, I.; Jiratova, K.; Pitchon, V.; Rynkowski, J.M. Effect of the preparation of supported gold particles on the catalytic activity in CO oxidation reaction. J. Mol. Catal. A Chem. 2005, 234, 187–197. [Google Scholar] [CrossRef]
  19. Rynkowski, J.M.; Dobrosz-Gómez, I. Ceria-Zirconia Supported Gold Catalysts. Annales Universitatis Mariae Curie-Skłodowska. Ann. UMCS Chem. 2009, 64, 197–217. [Google Scholar]
  20. Trovarelli, A.; Fornasiero, P. (Eds.) Catalysis by Ceria and Related Materials, 2nd ed.; Imperial College Press: London, UK, 2013. [Google Scholar]
  21. Dobrosz-Gómez, I.; Kocemba, I.; Rynkowski, J.M. Au/Ce1-xZrxO2 as effective catalysts for low-temperature CO oxidation. Appl. Catal. B Environ. 2008, 83, 240–255. [Google Scholar] [CrossRef]
  22. Dobrosz-Gómez, I.; Gómez-García, M.Á.; Szynkowska, M.I.; Kocemba, I.; Rynkowski, J.M. Surface, structural and morphological characterization of nanocrystalline ceria–zirconia mixed oxides upon thermal aging. Catal. Today 2012, 191, 142–145. [Google Scholar] [CrossRef]
  23. Dobrosz, I.; Gómez-García, M.Á.; Kocemba, I.; Maniukiewicz, W.; Rynkowski, J.M. Characterization of Au/Ce1-xZrxO2 catalysts in CO oxidation. Polish J. Environ. Studies 2005, 14, 231–234. [Google Scholar]
  24. Dobrosz-Gómez, I.; Kocemba, I.; Rynkowski, J.M. Factors influencing structure and catalytic activity of Au/Ce1-xZrxO2 catalysts in CO oxidation. Appl. Catal. B Environ. 2009, 88, 83–97. [Google Scholar] [CrossRef]
  25. Daturi, M.; Binet, C.; Lavalley, J.C.; Blanchard, G. Surface FTIR investigations on CexZr1−xO2 system. Surf. Interface. Anal. 2000, 30, 273–277. [Google Scholar] [CrossRef]
  26. Su, X.; Yang, X.F.; Huang, Y.Q.; Liu, B.; Zhang, T. Single-Atom Catalysis toward Efficient CO2 Conversion to CO and Formate Products. Acc. Chem. Res. 2019, 52, 656–664. [Google Scholar] [CrossRef]
  27. Wang, W.W.; Du, P.P.; Zou, S.H.; He, H.Y.; Wang, R.X.; Jin, Z.; Shi, S.; Huang, Y.Y.; Si, R.; Song, Q.S.; et al. Highly Dispersed Copper Oxide Clusters as Active Species in Copper-Ceria Catalyst for Preferential Oxidation of Carbon Monoxide. ACS Catal. 2015, 5, 2088–2099. [Google Scholar] [CrossRef]
  28. Esrafili, M.D.; Mousavian, P. A DFT Study on the Possibility of Using A Single Cu Atom Incorporated Nitrogen-Doped Graphene as A Promising and Highly Active Catalyst for Oxidation of CO. Int. J. Quant. Chem. 2019, 119, e25857. [Google Scholar] [CrossRef]
  29. Wang, H.; Shen, J.H.; Huang, J.F.; Xu, T.J.; Zhu, J.R.; Zhu, Y.H.; Li, C.Z. Atomically Dispersed Au Catalysts Supported on CeO2 Foam: Controllable Synthesis and CO Oxidation Reaction Mechanism. Nanoscale 2017, 9, 16817–16825. [Google Scholar] [CrossRef]
  30. Qiao, B.T.; Liu, J.X.; Wang, Y.G.; Lin, Q.Q.; Liu, X.Y.; Wang, A.Q.; Li, J.; Zhang, T.; Liu, J.Y. Highly Efficient Catalysis of Preferential Oxidation of CO in H2-Rich Stream by Gold Single-Atom Catalysts. ACS Catal. 2015, 5, 6249–6254. [Google Scholar] [CrossRef]
  31. Al Soubaihi, R.M.; Saoud, K.M.; Dutta, J. Critical Review of Low-Temperature CO Oxidation and Hysteresis Phenomenon on Heterogeneous Catalysts. Catalysts 2018, 8, 660. [Google Scholar] [CrossRef] [Green Version]
  32. Tang, Y.; Wang, Y.-G.; Li, J. Theoretical Investigations of Pt-CeO2 Single-Atom Catalyst for CO Oxidation. J. Phys. Chem. C. 2017, 121, 11281–11289. [Google Scholar] [CrossRef]
  33. Feng, Y.X.; Wang, Q.; Xiong, H.F.; Zhou, S.L.; Chen, X.; Hernandez, X.I.P.; Wang, Y.; Lin, S.; Datye, A.K.; Guo, H. Correlating DFT Calculations with CO Oxidation Reactivity on Ga-Doped Pt/CeO2 Single-Atom Catalysts. J. Phys. Chem. C 2018, 122, 22460–22468. [Google Scholar] [CrossRef]
  34. Kropp, T.; Lu, Z.; Li, Z.; Chin, Y.-H.C.; Mavrikakis, M. Anionic Single-Atom Catalysts for CO Oxidation: Support-Independent Activity at Low Temperatures. ACS Catal. 2019, 9, 1595–1604. [Google Scholar] [CrossRef]
  35. Yu, M.; Feng, Y.; Gao, L.; Lin, S. Phosphomolybdic Acid Supported Single-Metal-Atom Catalysis in CO Oxidation: First-Principles Calculations. Phys. Chem. Chem. Phys. 2018, 20, 20661–20668. [Google Scholar] [CrossRef]
  36. Wang, S.; Feng, Y.; Lin, S.; Guo, H. Phosphomolybdic Acid Supported Atomically Dispersed Transition Metal Atoms (M = Fe, Co, Ni, Cu, Ru, Rh, Pd, Ag, Os, Ir, Pt, and Au): Stable Single Atom Catalysts Studied by Density Functional Theory. RSC Adv. 2017, 7, 24925–24932. [Google Scholar] [CrossRef] [Green Version]
  37. Liu, X.; Sui, Y.H.; Duan, T.; Meng, C.G.; Han, Y. CO Oxidation Catalyzed by Pt-Embedded Graphene: A First-Principles Investigation. Phys. Chem. Chem. Phys. 2014, 16, 23584–23593. [Google Scholar] [CrossRef]
  38. Tang, Y.N.; Dai, X.Q.; Yang, Z.X.; Pan, L.J.; Chen, W.G.; Ma, D.W.; Lu, Z.S. Formation and Catalytic Activity of Pt Supported on Oxidized Graphene for the CO Oxidation Reaction. Phys. Chem. Chem. Phys. 2014, 16, 7887–7895. [Google Scholar] [CrossRef]
  39. Morgan, K.; Cole, K.J.; Goguet, A.; Hardacre, C.; Hutchings, G.J.; Maguire, N.; Shekhtman, S.O.; Taylor, S.H. TAP studies of CO oxidation over CuMnOX and Au/CuMnOX catalysts. J. Catal. 2010, 276, 38–48. [Google Scholar] [CrossRef]
  40. Schlexer, P.; Widmann, D.; Behm, R.J.; Pacchioni, G. CO Oxidation on a Au/TiO2 Nanoparticle Catalyst via the Au-Assisted Mars–van Krevelen Mechanism. ACS Catal. 2018, 8, 6513–6525. [Google Scholar] [CrossRef]
  41. Saqlain, M.A.; Hussain, A.; Siddiq, M.; Leitao, A.A. A DFT+U study of the Mars van Krevelen Mechanism of CO Oxidation on Au/TiO2 catalysts. Appl. Catal. A 2016, 519, 27–33. [Google Scholar] [CrossRef]
  42. Duan, Z.; Henkelman, G. CO Oxidation at the Au/TiO2 Boundary: The Role of the Au/Ti5c Site. ACS Catal. 2015, 5, 1589. [Google Scholar] [CrossRef]
  43. Widmann, D.; Behm, R.J. Activation of Molecular Oxygen and the Nature of the Active Oxygen Species for CO Oxidation on Oxide Supported Au Catalysts. Acc. Chem. Res. 2014, 47, 740–749. [Google Scholar] [CrossRef]
  44. Dobrosz, I.; Kocemba, I.; Rynkowski, J.M. The Optimisation of Au/Ce0.75Zr0.25O2 Catalyst Preparation for CO Oxidation. Pol. J. Environ. Studies 2006, 15, 44. [Google Scholar]
  45. Dobrosz-Gómez, I.; Kocemba, I.; Rynkowski, J.M. Carbon Monoxide Oxidation over Au/Ce1-xZrxO2 Catalysts: Effects of Moisture Content in the Reactant Gas and Catalyst Pretreatment. Catal. Lett. 2009, 128, 297–306. [Google Scholar] [CrossRef]
  46. Dobrosz-Gómez, I.; Gómez-García, M.Á.; Rynkowski, J.M. The Role of Au-Support Interactions in Creation of Catalytic Performance of Au/Ce0.75Zr0.25O2 in CO Oxidation. Pol. J. Environ. Studies 2009, 18, 587–591. [Google Scholar]
  47. Dobrosz-Gómez, I.; Gómez-García, M.Á.; Rynkowski, J.M. CO Oxidation over Au/CeO2–ZrO2 Catalysts: The Effect of the Support Composition of the Au Support Interaction. Kinet. Catal. 2010, 51, 823–827. [Google Scholar] [CrossRef]
  48. Gómez-García, M.Á.; Gómez-Mendoza, N.A.; Dobrosz-Gómez, I.; Gil Pavas, E.; Rynkowski, J.M. Temperature-Scanning Method for the Kinetic Studies of CO Oxidation over Ceria–Zirconia Supported Gold Catalysts. Chem. Eng. J. 2015, 282, 20–28. [Google Scholar] [CrossRef]
  49. Force, C.; Belzunegui, J.P.; Sanz, J.; Martinez-Arias, A.; Soria, J. Influence of precursor salt on metal particle formation in Rh/CeO2 catalysts. J. Catal. 2001, 197, 192–199. [Google Scholar] [CrossRef]
  50. Ivanova, S.; Pitchon, V.; Zimmermann, Y.; Petit, C. Preparation of alumina supported gold catalysts: Influence of washing procedures, mechanism of particles size growth. Appl. Catal. A Gen. 2006, 298, 57–64. [Google Scholar] [CrossRef]
  51. Schulz, A.; Hargittai, M. Structural variations and bonding in gold halides: A quantum chemical study of monometric and dimetric gold monohalide and gold trihalide molecules, AuX, Au2X2, AuX3, and Au2X6 (X = F, Cl, Br, I). Chem. Eur. J. 2001, 7, 3657–3670. [Google Scholar] [CrossRef]
  52. Arena, F.; Famulari, P.; Trunfio, G.; Bonura, G.; Frusteri, F.; Spadaro, L. Physicochemical properties and reactivity of Au/CeO2 catalysts in total and selective oxidation of CO. Appl. Catal. B Environ. 2006, 66, 81–91. [Google Scholar] [CrossRef]
  53. Fu, L.; Wu, N.Q.; Yang, J.H.; Qu, F.; Johnson, D.L.; Kung, M.C.; Kung, H.H.; Dravid, V.P. Direct evidence of oxidized gold on supported gold catalysts. J. Phys. Chem. Lett. B 2005, 109, 3704–3706. [Google Scholar] [CrossRef]
  54. Cao, A.; Lu, R.; Veser, G. Stabilizing Metal Nanoparticles for Heterogeneous Catalysis. Phys. Chem. Chem. Phys. 2010, 12, 13499–13510. [Google Scholar] [CrossRef]
  55. Cuenya, B.R. Synthesis and Catalytic Properties of Metal Nanoparticles: Size, Shape, Support, Composition, and Oxidation State Effects. Thin Solid Films 2010, 518, 3127–3150. [Google Scholar] [CrossRef]
  56. Jóźwiak, W.K.; Kaczmarek, E.; Ignaczak, W. Activity requirements of gold catalysts Au/Fe2O3 in low-temperature CO oxidation. Pol. J. Environ. Studies 2005, 14, 127. [Google Scholar]
  57. Moreau, F.; Bond, G.C.; Taylor, A.O. Gold on Titania Catalysts for the Oxidation of Carbon Monoxide: Control of PH during Preparation with Various Gold Contents. J. Catal. 2005, 231, 105–114. [Google Scholar] [CrossRef]
  58. Zanella, R.; Delannoy, L.; Louis, C. Mechanism of Deposition of Gold Precursors onto TiO2 during the Preparation by Cation Adsorption and Deposition–Precipitation with NaOH and Urea. Appl. Catal. A 2005, 291, 62–72. [Google Scholar] [CrossRef]
  59. Grzybowska-Świerkosz, B. Nano-Au/oxide Support Catalysts in Oxidation Reactions: Provenance of Active Oxygen Species. Catal. Today 2006, 112, 3–7. [Google Scholar] [CrossRef]
  60. Guzmán, J.; Carrettin, S.; Corma, A. Spectroscopic evidence for the supply of reactive oxygen during CO oxidation catalyzed by gold supported on nanocrystalline CeO2. J. Am. Chem. Soc. 2005, 127, 3286–3287. [Google Scholar] [CrossRef]
  61. Cunningham, D.A.H.; Vogel, W.; Haruta, M. Structural analysis of Au/TiO2 catalysts by debye function analysis. Catal. Lett. 1999, 63, 43–47. [Google Scholar] [CrossRef]
  62. Olea, M.; Tada, M.; Iwasawa, Y. TAP study on carbon monoxide oxidation over supported gold catalysts Au/Ti(OH)*4 and Au/Fe(OH)*3: Moisture effect. J. Catal. 2007, 248, 60–67. [Google Scholar] [CrossRef]
  63. Costello, C.K.; Yang, J.H.; Law, H.Y.; Wang, Y.; Lin, J.-N.; Marks, L.D.; Kung, M.C.; Kung, H.H. On the potential role of hydroxyl groups in CO oxidation over Au/Al2O3. Appl. Catal. A. Gen. 2003, 243, 15–24. [Google Scholar] [CrossRef]
  64. Ruszel, M.; Grzybowska, B.; Samson, K.; Gressel, I.; Klisinska, A. MIICr2O4-spinels as supports for Au nanoparticles in oxidation of CO. Catal. Today 2006, 112, 126–129. [Google Scholar] [CrossRef]
  65. Kang, Y.-M.; Wan, B.-Z. Gold and iron supported on Y-type zeolite for carbon monoxide oxidation. Catal. Today 1997, 35, 379–392. [Google Scholar] [CrossRef]
  66. Visco, A.M.; Neri, F.; Neri, G.; Donato, A.; Milone, C.; Galvagno, S.X. X-ray photoelectron spectroscopy of Au/Fe2O3 catalysts. Phys. Chem. Chem. Phys. 1999, 1, 2869–2873. [Google Scholar] [CrossRef]
  67. Pillai, U.R.; Deevi, S. Highly active gold-ceria catalyst for the room temperature oxidation of carbon monoxide. Appl. Catal. A Gen. 2006, 299, 266–273. [Google Scholar] [CrossRef]
  68. Abd El-Moemen, A.; Karpenko, A.; Denkwitz, Y.; Behm, R.J. Activity, stability and deactivation behavior of Au/CeO2 catalysts in the water gas shift reaction at increased reaction temperature (300 °C). J. Power Sources 2009, 190, 64–75. [Google Scholar] [CrossRef]
  69. Widmann, D.; Leppelt, R.; Behm, R.J. Activation of a Au/CeO2 catalyst for the CO oxidation reaction by surface oxygen removal/oxygen vacancy formation. J. Catal. 2007, 251, 437–442. [Google Scholar] [CrossRef]
  70. Wang, A.Q.; Hsieh, Y.P.; Chen, Y.F.; Mou, C.Y. Au–Ag alloy nanoparticle as catalyst for CO oxidation: Effect of Si/Al ratio of mesoporous support. J. Catal. 2006, 237, 197–206. [Google Scholar] [CrossRef]
  71. Qian, K.; Huang, W.; Fang, J.; Lv, S.; He, B.; Jiang, Z.; Wei, S. Low-temperature CO oxidation over Au/ZnO/SiO2 catalysts: Some mechanism insights. J. Catal. 2008, 255, 269–278. [Google Scholar] [CrossRef]
  72. del Río, E.; Collins, S.E.; Aguirre, A.; Chen, X.; Delgado, J.J.; Calvino, J.J.; Bernal, S. Reversible deactivation of a Au/Ce0.62Zr0.38O2 catalyst in CO oxidation: A systematic study of CO2-triggered carbonate inhibition. J. Catal. 2014, 316, 210–218. [Google Scholar] [CrossRef]
  73. Provendier, H.; Petit, C.; Schmitt, J.L.; Kiennemann, A.; Chaumont, C. Characterization of the solid solution La(Ni,Fe)O3 prepared via sol-gel related method using propionic acid. J. Mat. Sci. 1999, 34, 4121–4127. [Google Scholar] [CrossRef]
  74. Johnson, M.F.L.; Mooi, J. Cerium dioxide crystallite sizes by temperature-programmed reduction. J. Catal. 1987, 103, 502. [Google Scholar] [CrossRef]
  75. Johnson, M.F.L.; Mooi, J. Cerium dioxide crystallite sizes by temperature-programmed reduction. J. Catal. 1993, 140, 612. [Google Scholar] [CrossRef]
Figure 1. Characterization of (2 wt. % Au)/Ce1−xZrxO2 catalysts washed with different agents; below symbols: real gold-loading (wt. %); above symbols: Brunauer–Emmett–Teller (BET) -specific surface area (m2/g).
Figure 1. Characterization of (2 wt. % Au)/Ce1−xZrxO2 catalysts washed with different agents; below symbols: real gold-loading (wt. %); above symbols: Brunauer–Emmett–Teller (BET) -specific surface area (m2/g).
Catalysts 10 01312 g001
Figure 2. The back-scattered electron (BSE) photomicrographs, EDS maps and characteristic X-ray spectra of (2 wt. % Au)/C0.75Zr0.25O2 catalysts washed with different agents: (a) without washing (dried at 120 °C), (b) without washing (calcined at 300 °C), (c) washed with warm water (calcined at 300 °C), (d) washed with ammonia (calcined at 300 °C). (copyright permission number 4932711312351) [24].
Figure 2. The back-scattered electron (BSE) photomicrographs, EDS maps and characteristic X-ray spectra of (2 wt. % Au)/C0.75Zr0.25O2 catalysts washed with different agents: (a) without washing (dried at 120 °C), (b) without washing (calcined at 300 °C), (c) washed with warm water (calcined at 300 °C), (d) washed with ammonia (calcined at 300 °C). (copyright permission number 4932711312351) [24].
Catalysts 10 01312 g002aCatalysts 10 01312 g002b
Figure 3. The representative HRTEM images and the histograms of Au particle size distribution of Au supported on CeO2, Ce0.75Zr0.25O2 and ZrO2 washed with different agents: (a) prepared without washing, (b) washed with warm water, (c) washed with ammonia (a and c—copyright permission number 4932711312351) [24].
Figure 3. The representative HRTEM images and the histograms of Au particle size distribution of Au supported on CeO2, Ce0.75Zr0.25O2 and ZrO2 washed with different agents: (a) prepared without washing, (b) washed with warm water, (c) washed with ammonia (a and c—copyright permission number 4932711312351) [24].
Catalysts 10 01312 g003aCatalysts 10 01312 g003b
Figure 4. The resume of values of time-of-flight secondary ion mass spectrometry (ToF-SIMS)-selected secondary ions intensity ratios for some (2 wt. % Au)/Ce1−xZrxO2 catalysts prepared without washing and washed with ammonia.
Figure 4. The resume of values of time-of-flight secondary ion mass spectrometry (ToF-SIMS)-selected secondary ions intensity ratios for some (2 wt. % Au)/Ce1−xZrxO2 catalysts prepared without washing and washed with ammonia.
Catalysts 10 01312 g004
Figure 5. The back-scattered electron (BSE) photomicrographs, EDS maps and characteristic X-ray spectra of: (a) (2 wt. % Au)/CeO2, (b) (2 wt. % Au)/Ce0.75Zr0.25O2 and (c) (2 wt. % Au)/ZrO2 catalysts washed with ammonia (copyright permission number 4932711058170) [21].
Figure 5. The back-scattered electron (BSE) photomicrographs, EDS maps and characteristic X-ray spectra of: (a) (2 wt. % Au)/CeO2, (b) (2 wt. % Au)/Ce0.75Zr0.25O2 and (c) (2 wt. % Au)/ZrO2 catalysts washed with ammonia (copyright permission number 4932711058170) [21].
Catalysts 10 01312 g005aCatalysts 10 01312 g005b
Figure 6. X-ray diffraction patterns of (a) support (CeO2) and (2 wt. % Au)/CeO2 prepared without washing and washed with ammonia; (b) supports (Ce1−xZrxO2) and (2 wt. % Au)/Ce1−xZrxO2 catalysts washed with ammonia; (Δ) fluorite-type structure (cubic) of CeO2, (*) monoclinic-type structure of ZrO2, (♦) tetragonal-type structure of ZrO2.
Figure 6. X-ray diffraction patterns of (a) support (CeO2) and (2 wt. % Au)/CeO2 prepared without washing and washed with ammonia; (b) supports (Ce1−xZrxO2) and (2 wt. % Au)/Ce1−xZrxO2 catalysts washed with ammonia; (Δ) fluorite-type structure (cubic) of CeO2, (*) monoclinic-type structure of ZrO2, (♦) tetragonal-type structure of ZrO2.
Catalysts 10 01312 g006aCatalysts 10 01312 g006b
Figure 7. The effect of the washing procedure on CO oxidation activity of (2 wt. % Au)/Ce1−xZrxO2 catalysts: (a) prepared without washing; (b) washed with warm water; (c) washed with ammonia.
Figure 7. The effect of the washing procedure on CO oxidation activity of (2 wt. % Au)/Ce1−xZrxO2 catalysts: (a) prepared without washing; (b) washed with warm water; (c) washed with ammonia.
Catalysts 10 01312 g007
Figure 8. Temperature-programmed reduction (TPR-H2) (a) and TPR-CO (b) profiles of Ce0.75Zr0.25O2 support and (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts.
Figure 8. Temperature-programmed reduction (TPR-H2) (a) and TPR-CO (b) profiles of Ce0.75Zr0.25O2 support and (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts.
Catalysts 10 01312 g008aCatalysts 10 01312 g008b
Figure 9. The correlation between the hydrogen consumption, the ratio of oxygen atoms in the capping peak to the total cerium atoms and the activity of (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts prepared using different washing procedures.
Figure 9. The correlation between the hydrogen consumption, the ratio of oxygen atoms in the capping peak to the total cerium atoms and the activity of (2 wt. % Au)/Ce0.75Zr0.25O2 catalysts prepared using different washing procedures.
Catalysts 10 01312 g009
Figure 10. The effect of Ce1−xZrxO2 composition on the activity of (2 wt. % Au)/Ce1−xZrxO2 catalysts in CO oxidation (any activation procedure was used before the activity tests): (a) supports (for comparison); (b) prepared without washing; (c) washed with water; (d) washed with ammonia.
Figure 10. The effect of Ce1−xZrxO2 composition on the activity of (2 wt. % Au)/Ce1−xZrxO2 catalysts in CO oxidation (any activation procedure was used before the activity tests): (a) supports (for comparison); (b) prepared without washing; (c) washed with water; (d) washed with ammonia.
Catalysts 10 01312 g010
Figure 11. The influence of the pretreatment atmosphere on CO conversion at 30 °C on (2 wt. % Au)/Ce1−xZrxO2 catalysts washed with ammonia (CO:O2:He = 1.7:3.4:balance; W/F = 0.12 g s cm−3).
Figure 11. The influence of the pretreatment atmosphere on CO conversion at 30 °C on (2 wt. % Au)/Ce1−xZrxO2 catalysts washed with ammonia (CO:O2:He = 1.7:3.4:balance; W/F = 0.12 g s cm−3).
Catalysts 10 01312 g011
Figure 12. The effect of the pretreatment atmosphere on the reducibility of (2 wt. % Au)/Ce0.75Zr0.25O2 catalyst washed with ammonia.
Figure 12. The effect of the pretreatment atmosphere on the reducibility of (2 wt. % Au)/Ce0.75Zr0.25O2 catalyst washed with ammonia.
Catalysts 10 01312 g012
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dobrosz-Gómez, I.; Gómez-García, M.-Á.; Rynkowski, J.M. The Origin of Au/Ce1-xZrxO2 Catalyst’s Active Sites in Low-Temperature CO Oxidation. Catalysts 2020, 10, 1312. https://doi.org/10.3390/catal10111312

AMA Style

Dobrosz-Gómez I, Gómez-García M-Á, Rynkowski JM. The Origin of Au/Ce1-xZrxO2 Catalyst’s Active Sites in Low-Temperature CO Oxidation. Catalysts. 2020; 10(11):1312. https://doi.org/10.3390/catal10111312

Chicago/Turabian Style

Dobrosz-Gómez, Izabela, Miguel-Ángel Gómez-García, and Jacek Michał Rynkowski. 2020. "The Origin of Au/Ce1-xZrxO2 Catalyst’s Active Sites in Low-Temperature CO Oxidation" Catalysts 10, no. 11: 1312. https://doi.org/10.3390/catal10111312

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