Next Article in Journal
Facile Synthesis of High Performance Iron Oxide/Carbon Nanocatalysts Derived from the Calcination of Ferrocenium for the Decomposition of Methylene Blue
Next Article in Special Issue
Electric Field Promoted Complete Oxidation of Benzene over PdCexCoy Catalysts at Low Temperature
Previous Article in Journal
Kinetic Monte-Carlo Simulation of Methane Steam Reforming over a Nickel Surface
Previous Article in Special Issue
Structure and Catalytic Behavior of Alumina Supported Bimetallic Au-Rh Nanoparticles in the Reduction of NO by CO
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

{CeO2/Bi2Mo1−xRuxO6} and {Au/Bi2Mo1−xRuxO6} Catalysts for Low-Temperature CO Oxidation

by
Edson Edain González
1,
Ricardo Rangel
1,*,
Javier Lara
1,
Pascual Bartolo-Pérez
2,
Juan José Alvarado-Gil
2,
Donald Homero Galván
3 and
Rafael García
4
1
División de Estudios de Posgrado de la Facultad de Ingeniería Química, Universidad Michoacana de San Nicolás de Hidalgo, Gral. Francisco J. Múgica S/N, Ciudad Universitaria. Z. C. 58030 Morelia, Michoacán, Mexico
2
Departamento de Física Aplicada, CINVESTAV, Unidad Mérida, Antigua Carretera a Progreso Km. 6, Cordemex. Z. C. 97310 Mérida, Yucatán, Mexico
3
Centro de Nanociencias y Nanotecnología, Universidad Nacional Autónoma de México, Z. C. 22800 Ensenada, B.C., Mexico
4
Departamento de Investigación en Física, Universidad de Sonora, 83000 Hermosillo, Sonora, Mexico
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(11), 947; https://doi.org/10.3390/catal9110947
Submission received: 8 October 2019 / Revised: 2 November 2019 / Accepted: 6 November 2019 / Published: 12 November 2019
(This article belongs to the Special Issue Catalysis for the Removal of Gas-Phase Pollutants)

Abstract

:
Nowadays, one of the most important challenges that humanity faces is to find alternative ways of reducing pollutant emissions. CeO2/Bi2Mo1−xRuxO6 and Au/Bi2Mo1−xRuxO6 catalysts were prepared to efficiently transform carbon monoxide (CO) to carbon dioxide (CO2) at low temperatures. The systems were prepared in a two-step process. First, Bi2Mo1−xRuxO6 supports were synthesized through the hydrothermal procedure under microwave heating. Then, CeO2 was deposited on Bi2Mo1−xRuxO6 using the wet impregnation method, while the incipient impregnation method was selected to deposit gold nanoparticles. The CeO2/Bi2Mo1−xRuxO6 and Au/Bi2Mo1−xRuxO6 catalysts were characterized using SEM microscopy and XRD. Furthermore, energy-dispersive X-ray spectroscopy (EDS), X-ray photoelectron spectroscopy (XPS), and Raman spectroscopy were used. Tests were carried out for the supported catalysts in CO oxidation, and high conversion values, nearing 100%, was observed in a temperature range of 100 to 250 °C. The results showed that the best system was the Au/Bi2Mo0.95Ru0.05O6 catalyst, with CO oxidation starting at 50 °C and reaching 100% conversion at 186 °C.

Graphical Abstract

1. Introduction

The drastic growth in atmospheric pollution is the product of industrial development as well as the increasing use of automobiles for transportation, which in turn maintains a growing demand for fossil fuels. It is an important and still pending task to decrease the atmospheric pollutants coming from these processes, including COx, SOx, and NOx, which significantly contribute to the phenomena of acid rain, global warming, and destruction of the ozone layer [1]. Carbon monoxide (CO) is considered highly poisonous, and it causes severe health issues and irreversible environmental damage [2]. In this context, the development of materials with high activity for removing CO becomes necessary. Traditional supported noble metal catalysts are examples of materials with excellent catalytic properties to promote CO oxidation at low temperatures [3,4].
Bi2MoO6 is recognized as a good catalyst, and it is useful in selective oxidation processes [5]. The most common methods to synthesize it include coprecipitation [6,7], followed by calcination processes or through solid-state reactions [8] at temperatures above 1000 °C. Alternatively, Bi2MoO6 can also be obtained by sol–gel synthesis [9] or spray-drying methods [10]. The catalytic properties of bismuth molybdate are well known. For instance, in propane transformation reactions, it is considered an essential component [6,11], even though some of its phases are not active, as has been reported by Ueda et al. [12]. Schuh et al. [13] investigated the selectivity of several bismuth molybdates in propylene oxidation reaction using α-Bi2Mo3O12, β-Bi2Mo2O9, and γ-Bi2MoO6 phases. When testing the γ-Bi2MoO6 catalyst containing an excess of stoichiometric bismuth/molybdenum in a 3:1 ratio, the authors obtained results for propylene conversion of up to 7.3%. However, the selectivity toward acrolein formation was very low, and mainly COx was formed. Compounds with a Bi/Mo ratio of 3:1 was found to produce a α-Bi2Mo3O12 compound exhibiting acrolein selectivity in the range of 36%–56%, providing a 60%–63% propylene conversion. Brazdil et al. [14] reported on the synthesis of bismuth–cerium molybdate compounds with a BixCe2−xMo3O12 formulation. They studied the selectivity as a function of cerium content, to prove its effectiveness in reactions for catalytic propylene ammoxidation. Their results allowed them to establish active sites as a function of the bismuth content for propylene ammoxidation. Other investigations have been conducted using Bi2MoxW1−xO6 supported on Al2O3–SiO2, which were tested in carbon monoxide oxidation reaction to evaluate its activation temperature. The compounds were found to activate at 200 °C, reaching 100% conversion at about 350 °C [15]. Furthermore, there have been investigations based on the dehydrogenation of organic compounds using catalysts composed of mixed oxides containing molybdenum, bismuth, and vanadium. The studies examined their catalytic activity and determined how the bismuth–molybdate structure is modified when it is doped with a metal. In a temperature-programmed reoxidation experiment, they established that catalysts containing vanadium exhibited enhanced oxygen mobility in comparison to Bi2MoO6, showing an important 1-butene to 1,3-butadiene catalytic conversion (72%) in oxidative dehydrogenation reactions [16,17].
Ruthenium is not a widely used metal in oxidation reactions due to its low melting point as RuO2. However, its high reactivity toward carbon monoxide oxidation has become a point of interest in the field of catalysis. In this regard, studies have been conducted on ruthenium in relation to their particle size, which ranges from 2 to 6 nm, using polyvinylpyrrolidone (PVP) as the size-regulating agent. The results showed that, as Ru nanoparticles became smaller, they began to be more active and efficient in CO to CO2 conversion [18]. Another reaction in which ruthenium is used as a catalyst is in the heterogeneous oxidation of alcohols. To induce these reactions, ruthenium is obtained by coprecipitation processes and deposited on Al2O3 metal supports being activated in an oxygen-rich atmosphere [19]. Ruthenium nanoparticles deposited on cerium oxide has also been reported, with the studies concluding that a relatively large CeO2 surface area favors ruthenium particle dispersion, increasing its catalytic activity in CO oxidation reaction [19,20]. Unfortunately, RuO2 is not stable unless it is stabilized by the addition of some external element, which makes its use technically complicated. The key to the widespread use of RuO2 is therefore to find a way to achieve such stabilization.
In the last few decades, researchers have studied the catalytic properties of supported gold catalysts for CO oxidation. Au nanoparticles have been successfully tested in carbon monoxide oxidation reactions at low temperatures [21]. One of the first investigations carried out on the synthesis of gold nanoparticles was performed by Haruta [22], who reported the production of synthesized particles smaller than 10 nm and deposited on oxide supports for its use in hydrogen and carbon monoxide oxidation reactions. Other investigations examined the use of gold nanoparticles deposited on ceria nanotubes and obtained good results in catalytic oxidation reactions [23]. Such materials perform an important role in catalyzing different reactions at low temperatures. This feature is attributed to the presence of low-coordinate gold atoms on the surface of CeO2 nanoparticles [24]. CeO2 powder has been listed as one of the best supporting materials for catalysis when it is associated with gold nanoparticles as an active phase due to its relevant storage and oxygen-releasing capability [25]. An important feature attributed to CeO2 compounds, is the formation of vacancies within the crystalline structure, when doping is carried out because these could modify the entropy and increase its thermodynamic stability. These features are relevant as they determine the speed at which ions can move through their surface [26], providing bandgap narrowing, which in turn promotes the transference of electrons from the valence to the conduction band [21,25,27,28]. The oxidative capability of cerium oxide has been widely demonstrated through several studies on CO oxidation reactions, used either as an active phase [5] or as a support in noble metal catalysts [25]. Furthermore, specific studies on nanostructured CeO2 powders in the form of rods exposing predominantly (111) surfaces and cubes exposing (100) surfaces have been reported. In those studies have been reported that Pd deposited on the (111) plane of ceria rods, exhibited CO conversion at room temperature, while Pd on (100) ceria cubes showed similar CO conversion values at 60 °C. The differences found in CO conversion can be attributed to Pd redispersion, which is more pronounced in oxidative conditions for Pd/CeO2 rods [29].
The results already discussed related to CeO2, RuO2, Au nanoparticles, and Bi2MoO6 used as catalysts in CO oxidation reaction, supports the hypothesis that different prepared materials including oxides as CeO2/Bi2Mo1−xRuxO6 (OCMBR) and Au/Bi2Mo1−xRuxO6 (AuMBR), could oxidize efficiently CO to CO2. It is also important to elucidate the synergistic relationship between the active phases and the mixed oxide support to promote CO to CO2 conversion at low temperatures.

2. Results

2.1. Scanning Electron Microscopy (SEM)

The morphology of the MBR supports and OCMBR and AuMBR catalysts are shown in Figure 1a,b and Figure 1c–f, respectively. In Figure 1a,b, granular and small rounded structures can be observed. The 500 nm particles had a similar shape but different distributions. In the case of Bi2Mo0.95Ru0.05O6 (MBR1), there was formation of agglomerations, while the particles were smaller in Bi2Mo0.97Ru0.03O6 (MBR2). Figure 1c–f shows OCMBR and AuMBR catalysts consisting of rounded particles (< 70 nm) obtained as a result of using the PVP surfactant agent. In this case, the particles had an average size of approximately 100 nm.
Table 1 shows the chemical composition of the catalysts (Ce, Au, Bi, Mo, O, Ru). The chemical content was determined by energy-dispersive X-ray spectroscopy (EDS). The atomic concentration of the species corresponding to bismuth molybdate remained constant, with a high concentration of oxygen. It is important to mention that the quantification performed for Au and Ce showed values close to 1.5 atomic percent (at %) which is lower than the expected 2 (at %) content. This could be attributed to some loss during the deposition or washing processes. In the case of cerium, the values exceeded only 3 at %.

2.2. X-Ray Diffraction (XRD)

X-ray diffraction patterns of the synthesized compounds, including the doped ruthenium catalytic support and the catalysts after impregnation, are shown in Figure 2. The signal corresponding to the 2θ angle of 28.34° can be attributed to the (131) plane characteristic of the Bi2MoO6 gamma phase, which has an orthorhombic-like structure. Two other signals located at 32.51° (200) and 47.10° (260) were also found (Figure 2f).
The synthesis allowed correct ruthenium incorporation into the catalytic support structure [30]. This was verified through the signal displacement of the (131) plane toward greater angles (Figure 2a,b).
Regarding the OCMBR compounds, their X-ray diffraction pattern showed a signal located at 28.4° (2θ), as shown in Figure 2c,d. The signal can be attributed to the CeO2 (111) plane, in agreement with results previously reported by other authors [31]. For the Au/Bi2Mo0.95Ru0.05O6 (AuMBR1) catalyst, characteristic peaks of the gamma phase of bismuth molybdate (γ-Bi2MoO6) were found. Their main characteristic peak located at 28.3° (Figure 2e) can be attributed to the (131) crystalline plane, on which it obtained a crystal size of 37.9 and 37 nm for MBR1 and MBR2, respectively. These values can be obtained through the Scherrer equation:
D = K λ W cos θ
where K is a constant that depends on particle morphology (usually K = 1 for cubic or nearly cubic systems), λ is the Cu Kα radiation (nm), W is the full width at half maximum (rad), and θ is the diffraction angle (deg).

2.3. X-Ray Photoelectron Spectroscopy (XPS)

The XPS technique was used to detect the presence of different chemical species involved in our catalysts. The binding energy (BE) analysis for the chemical species of the support revealed signals at 530.57, 527.4, and 529.9 eV for oxygen (O1s). The peaks for Mo3d and Bi4f were also detected. The signals for Mo+6 were obtained at 235.66 and 232.63 eV [32]. In addition, the signals for Bi+3 were observed at 164.63 and 159.37 eV (Figure 3) [30,33]. An XPS signal, located at 280 eV, was found corresponding to Ru3d5/2. This is often confused with carbon signal, but a high-resolution analysis performed for this compound confirmed its presence (Figure 4). This BE signal matched Ru+4 [34] and was detected at 280.6 and 280.3 eV. These signals can be attributed to the presence of Ru–O bonds within the Bi2Mo1−xRuxO6 structure. This outcome agrees with the results discussed in the XRD section, where the ruthenium oxide phase was not found, similar to previous findings [35].
Figure 5 shows the XPS analysis for CeO2/Bi2Mo0.95Ru0.05O6 (OCMBR1) and CeO2/Bi2Mo0.97Ru0.03O6 (OCMBR2) supported catalysts. In addition to Bi, Ru, and O, the representative signal of cerium (Ce3d) with a BE of 886.08 eV, corresponding to the Ce–O interaction in CeO2, was established [36,37,38].
Figure 6 shows the XPS scans of AuMBR catalysts, where Bi, Ru, O, and Au signals were detected. The oxidation state for every element present was matched according to established standards. In addition, a high-resolution energy spectrum for gold was performed to obtain signals for the different oxidation states present. The results showed the presence of gold as Au+3 and Au0. The corresponding XPS curves of Au spectra of the Au/MBR catalysts are shown in Figure 7a–d. The characteristic signals for Au0 and Au+3 were found at the BE of 84 and 87.43 eV values, respectively, which were assigned to Au 4f7/2 and Au 4f5/2 signals. The 4f7/2 signal revealed that the Au was in a metallic valence, while the 4f5/2 signal was attributed to its ionic state [39,40]. The results of the deconvolution procedure to establish the amount of Au+3 and Au0 are summarized in Table 2.
To increase the presence of Au0 in both AuMBR1 and Au/Bi2Mo0.97Ru0.03O6 (AuMBR2) catalysts, they were subjected to a reduction process using H2 at 500 °C for 4 h. The results for Au+3 and Au0 are shown in Table 2 and also in Figure 7b,d [41,42,43].
After reducing each AuMBR catalyst with H2, the atomic quantification of each species was performed, as shown in Figure 7. Table 2 gives the results of Au0 content before and after carrying out the reduction process. It can be observed that, for the AuMBR1 catalyst, the Au+3 percentage was higher than that of Au0 before reduction. Once the reduction process was carried out, the Au0/Au+3 ratio increased, thus obtaining a higher amount of Au0. The corresponding analyses for the reduced samples—AuMBR1_R and AuMBR2_R—are shown in Figure 7b,d. As can be seen, the Au0 content increased by about 30 at %. The oxidative capacity of the catalysts was studied before and after the reduction procedure, and their corresponding values will be discussed later.

2.4. Raman Spectroscopy

Figure 8 shows the spectrum obtained for MBR supports, with their vibrational characteristic mode values located at 68.16, 131.86, 788.78, and 880.8 cm−1. Similar results have been reported by other authors, establishing that very strong vibrational modes are observed for Bi2MoO6 at 808 cm−1 as well as two minor peaks at 852 and 722 cm−1, corresponding to Mo–O stretches of the distorted MoO6 octahedra [44]. Additionally, two signals found at 356 and 283 cm−1, which can be ascribed to the Bi–O–Mo or Bi–O vibrational modes [34,35,45,46], were shown in the Raman spectra when a β-Bi2MoO6 phase was detected. The Ru characteristic band with Ru–O vibrational modes can be found at 310 cm−1 [36,41]. As can be seen from Figure 8, the ruthenium content promoted the shifting from 299.73 to 301.72 cm−1.

2.5. Catalytic Activity Tests

In this section, the results obtained for tests carried out on the capacity of the developed compounds for carbon monoxide oxidation as a function of temperature are presented. MBR1 and MBR2 supports showed an improvement in comparison to pristine Bi2MoO6 (MB), as previously reported [5]. MB showed its activation at 220 °C, reaching 98% CO oxidation at 380 °C. The MB support never reached 100% conversion, as can be seen in Figure 9, which means that the support decreases its oxidative capacity at a temperature above 300 °C. In comparison, MBR1 and MBR2 compounds started their activation under 150 °C and reached 98% CO oxidation at 220 °C, demonstrating an improvement due to ruthenium inclusion. The difference in conversion capacity between MB and MBR catalysts can be attributed to the atomic content of ruthenium in the MBR supports, in accordance with previously reported results [42].
Figure 10 shows the CO conversion behavior of AuMBR and OCMBR catalysts. The OCMBR1 catalyst was activated at 150 °C, reaching 100% conversion at 250 °C, while the OCMBR2 catalyst was activated at 175 °C and reached 100% CO conversion at 275 °C. The OCMBR catalysts were found to be more efficient in CO conversion than MBR catalytic supports, being activated at 95 °C and reaching 100% CO conversion at 280 °C. This is related to the oxidative capacity of ceria, which works synergistically with the Bi2Mo1−xRuxO6 compound [47]. Venkataswamy et al. [48] performed work on Mn–ceria-doped solid solutions used as catalysts in CO oxidation reactions at low temperatures. They observed that Mn–ceria catalysts were activated at 50 °C, reaching 100% conversion at a temperature higher than 320 °C. It is necessary to highlight that OCMBR catalysts, in spite of being a highly active compound in CO oxidation [21,49], were not the ones with the best results, as shown in the conversion graph.
As described in the section on XPS analysis, AuMBR catalysts showed a higher concentration of Au+3 in comparison to Au0. Based on the analysis, it can be inferred that the catalysts underwent a reduction with hydrogen to increase the Au0 concentration (see Table 2). The reduced catalysts were labeled as AuMBR_R. Figure 10 shows the results of CO conversion for the nonreduced catalysts and the reduced ones. It can be observed that the AuMBR1 and AuMBR2 catalysts were activated at 150 °C, reaching 100% conversion at 275 °C and 300 °C, respectively. In the case of the reduced catalysts—AuMBR1_R and AuMBR2_R—they were activated at 23 °C. The AuMBR1_R catalyst reached 100% CO conversion at 186 °C, while the AuMBR2_R catalyst reached it at 225 °C. There was substantial change in terms of oxidative capacity of the gold supported catalysts when they underwent a reduction process. This means that a higher Au0/Au+1 ratio, obtained through the reduction process, improves CO oxidation [18,21,50].

3. Discussion

Several studies have been carried out to understand how CO molecules interact with Au particles supported on ZnO [51]. According to the obtained results, the oxidation process starts with a process of CO adsorption taking place in CO/Au/ZnO complexes in specific sites formed by Au clusters following the Mars–van Krevelen (MvK) kinetic model [52]. AuMBR catalysts work similarly, carrying out the CO oxidation process through a MvK kinetic model in synergy with the oxidative capability of the MBR (Bi2Mo1−xRuxO6) support, which enhances the transport efficiency of electrons toward the surface, thereby improving the oxidative activity of Au particles. The formation of the complex CO/Au/Bi2Mo1−xRuxO6 is believed to happen in the same order as the MvK mechanism.
Regarding the OCMBR catalysts, the reaction mechanism follows a consecutive series of elementary steps that include a constant change in the cerium valence (Ce+3 ↔ Ce+4) due to changes in oxygen concentration occurring within the OCMBR structure. In CO oxidation reactions, the oxygen storage capability of CeO2 catalysts has been demonstrated. Oxygen located at the catalyst surface oxidize the CO molecules, while the introduced bimolecular oxygen decomposes according to the Langmuir-Hinshelwood mechanism to fill the vacancies created by the CO oxidation reaction [53].
Investigations with Au nanoparticles have reported obtaining compounds capable of being activated at 0 °C. However, they do not reach a high conversion rate until temperatures above 300 °C [17,43,54]. In the case of compounds in which Au acts as active phase (Au/TiO2, Au/NiO, Au/α-Fe2O3), catalysts show a disadvantage because the formation of water vapor at high temperatures increases to the point of reducing the oxidative activity of the Au nanoparticles, as shown by other similar reports [21,23,38,41]. An important challenge in the present research was to ensure that most of the dispersed gold particles on the Bi2Mo1−xRuxO6 surface supports could be found in the metallic oxidation state (Au0) to further increase its oxidative behavior. Our XPS analysis demonstrated that gold was found in Au+1 and Au0 oxidation states, with Au0 more predominant, providing our catalysts an outstanding oxidative capacity. This was demonstrated in the catalytic activity tests, where AuMBR1 reached 100% CO oxidation at 150° C. Furthermore, in our studies, metallic particles smaller than 10 nm in size were obtained, which is a determining feature to achieve the CO oxidation reaction, as previously discussed by Zhang et al. [55] and Kimling et al. [56]. The results here discussed are encouraging for continued work on the optimization of our systems.

4. Materials and Methods

4.1. Catalysts Preparation

MBR1 and MBR2 solid solutions were prepared starting with aqueous solutions of (NH4)6Mo7O24 (97%, Alyt), Bi(NO3)3 (99.99%, Alfa Aesar), and Ru3(CO)12 (99%, Alfa Aesar), which were mixed in a stoichiometric ratio. PVP was also added into the mixture as a capping agent to control particle size. The mixture was stirred for about two hours at 30 °C. Then, the solid solution was formed using hydrothermal treatment assisted by microwave heating at 150 °C for 1 h. The resulting powder was washed with 5 mL of isopropanol, dried at 120 °C for 24 h, and then heated at 500 °C for 3 h. The supported systems—AuMBR1 and AuMBR2—were obtained using HAuCl4 solution 0.0016 M as a gold precursor and urea 0.42 M, through the deposit–precipitation (DP) urea method [57]. The Au/catalyst ratio was calculated as 1% (w/w).
Regarding OCMBR1 and OCMBR2 catalysts, they were prepared via wet impregnation using a 0.23 M Ce(NO3)3 solution. The stoichiometric ratio of the active phase to the support was nearly 4% (w/w).
After performing the impregnation procedure, all of the catalysts were dried at 120 °C for 12 h and calcined at 550 °C for 3 h. Before the catalytic tests, AuMBR1 and AuMBR2 catalysts were reduced at 400 °C for 4 h in a 30 cm3min−1 hydrogen flow.

4.2. Characterization

SEM images were obtained using a JEOL JSM 5300 scanning electron microscope Noran Instruments (Japan), coupled with an energy-dispersive analyzer (EDAX) also provided by Noran Instruments (Japan) operating at 20 kV and 1 × 10−6 Torr. X-ray diffraction analyses were obtained in a Philips X’Pert analytical diffractometer (England) using Cu Kα radiation under conditions of 40 kV voltage and 30 mA current. Before the analysis, the catalysts were ground in an agate mortar and passed through a 100-mesh sieve (Tyler series). The XPS technique allowed the oxidation states to be determined, in which the main elements were found using a Thermo Scientific equipment with an X-ray source provided with an Al-Kα monochromator (MA, USA). The analysis area was 400 μm2. Before starting the analysis, the surface was eroded with low-energy Ar ions for 15 s to eliminate possible surface impurities. Raman spectroscopy analyses were performed on a Witec Alpha 300 device (Germany) using a laser as the illumination source (633 nm) provided with a charge-coupled device (CCD) detector for the measurement of scattered Raman emission.

4.3. Catalytic Assessment

The CO oxidation reaction was carried out in a packed microflow quartz reactor, 8 mm diameter, under atmospheric pressure in a range of 20–350 °C. The CO to O2 ratio was adjusted to CO/O2 = 1 with a rate of 60 mL/min (2%, CO, 2% O2, 96% Ar). The catalyst weight was 100 mg. Before the tests, catalysts were pretreated (activated) at 300° C with a mixture of H2/N2 gases in a 2%/90% v/v ratio, for 1 h. The reaction products were analyzed using thermal conductivity gas chromatography provided with a carboxen-1000 column (1/8 inch diameter and 4 m long) using a thermal conductivity detector (TCD) to determine the reaction products, as described in Figure 11.

5. Conclusions

In the present study, the oxidative capability of CeO2/Bi2Mo1−xRuxO6 and Au/Bi2Mo1−xRuxO6 catalysts were demonstrated. It was found that the ruthenium inclusion as Bi2Mo1−xRuxO6 improved the oxidative behavior compared to Bi2MoO6. The Bi2Mo1−xRuxO6 compound achieved 100% CO conversion at 260 °C, which can be considered an improvement compared to the Bi2MoO6 compound, which achieved 95% CO conversion at 375 °C. Regarding the supported catalysts, a synergistic effect was observed between the active phase and the catalytic support. The CeO2/Bi2Mo1−xRuxO6 system showed an improvement, with OCMBR1 starting CO conversion at a temperature of 120 °C and reaching 100% conversion to CO2 at 225 °C. With the Au/Bi2Mo1−xRuxO6 catalytic systems, it started oxidizing CO at a temperature as low as 23 °C, reaching 100% CO conversion at a temperature of 180 °C. The Au/Bi2Mo0.97Ru0.03O6 compound was found to be the best catalyst to oxidize CO at low temperatures. Further optimization studies are needed regarding the stability and oxidative capability of Au and CeO2 supported systems.

Author Contributions

Conceptualization, R.R. and E.E.G.; methodology, E.E.G.; validation, E.E.G. and. R.R.; formal analysis, R.R., E.E.G., J.L., P.B.-P., J.J.A.-G., D.H.G., and R.G.; investigation, R.R. and E.E.G.; resources, R.R., P.B.-P., J.J.A.-G., and R.G.; data curation, E.E.G.; writing—original draft preparation, R.R. and E.E.G.; writing—review and editing, R.R., E.E.G., J.L., P.B.-P., J.J.A.-G., and D.H.G.; funding acquisition, R.R.

Funding

This research received no external funding.

Acknowledgments

E. González acknowledges the scholarship provided by CONACyT. R. Rangel acknowledges the CIC-UMSNH project 2019. The authors thank W. Cauich for technical support in the XPS and SEM analyses, D. Aguilar for DRX analysis, and J. Bante for Raman spectroscopy analysis. The authors thank LANBIO-CINVESTAV and FOMIX-Yucatán 2008-108160 Conacyt LAB-2009-01-123913, 292692, 294643 projects.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tit, N.; Said, K.; Mahmoud, N.M.; Kouser, S.; Yamani, Z.H. Ab-initio investigation of adsorption of CO and CO2 molecules on graphene: Role of intrinsic defects on gas sensing. Appl. Surf. Sci. 2017, 394, 219–230. [Google Scholar] [CrossRef]
  2. Zhao, Y.; Hu, J.; Tan, Z.; Liu, T.; Weilin, Z.; Li, X.; Huang, C.; Wang, S.; Huang, Z.; Ma, W. Ambient carbon monoxide and increased risk of daily hospital outpatient visits for respiratory diseases in Dougguan, China. Sci. Total Environ. 2019, 668, 254–260. [Google Scholar] [CrossRef] [PubMed]
  3. Ivanov, K.I.; Kolentsova, E.N.; Dimitrov, D.Y.; Avdeev, G.; Tabakova, T. Alumina supported copper-manganese catalysts for combustion of exhaust gases: Catalysts characterization. In Proceedings of the XIII International Conference on Chemical Engineering and Applications, Hong Kong, 27–28 August 2015. [Google Scholar]
  4. Dey, S.; Dhal, G.C. Materials progress in the control of CO and CO2 emission at ambient conditions: An overview. Mater. Sci. Energy Technol. 2019, 2, 607–623. [Google Scholar] [CrossRef]
  5. Rangel, R.; Maya-Yescas, R.; García, R. Novel [Ce1−xLaxO2, La2−yCeyO3]/Bi2Mo0.9W0.1O6 catalysts for CO oxidation at low temperature. Catal. Sci. Technol. 2012, 2, 639–642. [Google Scholar] [CrossRef]
  6. Keulks, G.W.; Hall, J.L.; Suzuki, C.D. The catalytic oxidation of propylene: IV. Preparation and characterization of a-bismuth molybdate. J. Catal. 1974, 34, 79–97. [Google Scholar] [CrossRef]
  7. Trifiro, F.; Hoser, H.; Scarle, R.D. Relationship between structure and activity of mixed oxide as oxidation catalysts. I. Preparation and solid-state reactions of Bi-molybdates. J. Catal. 1972, 25, 12–24. [Google Scholar] [CrossRef]
  8. Rastogi, R.P.; Singh, A.; Shukla, C.S. Kinetics and mechanism of solid-state reaction between bismuth(III) oxide and molybdenum(VI) oxide. J. Solid State Chem. 1982, 42, 136–148. [Google Scholar] [CrossRef]
  9. Le, M.T.; Bac, L.H.; van Driessche, I.; Hoste, S.; van Well, W.J.M. The synergy effect between gamma and beta phase of bismuth molybdate catalysts: Is there any relation between conductivity and catalytic activity? Catal. Today 2008, 131, 566–571. [Google Scholar] [CrossRef]
  10. Le, M.T.; van Craenenbroeck, J.; van Driessche, I.; Hoste, S. Bismuth molybdate catalysts synthesized using spray drying for the selective oxidation of propylene. Appl. Catal. A Gen. 2003, 249, 355–364. [Google Scholar] [CrossRef]
  11. Carrazán, S.R.G.; Martín, C.; Mateos, R.; Rives, V. Influence of the active phase structure Bi-Mo-Ti-O in the selective oxidation of propene. Catal. Today 2006, 112, 121–125. [Google Scholar] [CrossRef]
  12. Moro-oka, Y.; Ueda, W. Multicomponent bismuth molybdate catalyst: A highly functionalized catalyst system for the selective oxidation of olefin. Adv. Catal. 1994, 40, 233–273. [Google Scholar]
  13. Schuh, K.; Kleist, W.; Høj, M.; Trouillet, V.; Beato, P.; Jensen, A.D.; Patzke, G.R.; Grunwaldt, J.D. Selective oxidation of propylene to acrolein by hydrothermally synthesized bismuth molybdates. Appl. Catal. A Gen. 2014, 482, 145–156. [Google Scholar] [CrossRef]
  14. Brazdil, J.F.; Toft, M.A.; Lin, S.S.Y.; McKenna, S.T.; Zajac, G.; Kaduk, J.A.; Golab, J.T. Characterization of bismuth-cerium-molybdate selective propylene ammoxidation catalysts. Appl. Catal. A Gen. 2015, 495, 115–123. [Google Scholar] [CrossRef]
  15. Rangel, R.; López, J.L.C.; Espino, J.; Núñez-González, R.; Bartolo-Pérez, P.; Gómez-Cortés, A.; Díaz, G. Estudio del efecto de diferentes soportes mixtos en la actividad catalitica y las caracteristicas estructurales de catalizadores de Bi2MoxW1−xO6. Rev. Int. Contam. Ambient. 2014, 30, 841–848. [Google Scholar]
  16. Park, J.H.; Shin, C.H. Influence of the catalyst composition in the oxidative dehydrogenation of 1-butene over BiVxMo1−x oxide catalysts. Appl. Catal. A Gen. 2015, 495, 1–7. [Google Scholar] [CrossRef]
  17. Joo, S.H.; Park, J.Y.; Renzas, J.R.; Butcher, D.R.; Huang, W.; Somorjai, G.A. Size effect of ruthenium nanoparticles in catalytic carbon monoxide oxidation. Nano Lett. 2010, 10, 2709–2713. [Google Scholar] [CrossRef] [PubMed]
  18. Yamaguchi, K.; Mizuno, N. Supported ruthenium catalyst for the heterogeneous oxidation of alcohols with molecular oxygen. Angew. Chem. Int. Ed. 2002, 114, 4538–4542. [Google Scholar] [CrossRef]
  19. Satsuma, A.; Yanagihara, M.; Ohyama, J.; Shimizu, K. Oxidation of CO over Ru/Ceria prepared by self-dispersion of Ru metal powder into nano-sized particle. Catal. Today 2013, 201, 62–67. [Google Scholar] [CrossRef]
  20. Gaálová, J.; Barbier, J.; Rossignol, S. Ruthenium versus platinum on cerium materials in wet air oxidation of acetic acid. J. Hazard. Mater. 2010, 181, 633–639. [Google Scholar] [CrossRef] [PubMed]
  21. Kim, H.Y.; Lee, H.M.; Henkelman, G. CO oxidation mechanism on CeO2-supported Au nanoparticles. J. Am. Chem. Soc. 2011, 134, 1560–1570. [Google Scholar] [CrossRef] [PubMed]
  22. 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]
  23. Zhang, R.; Lu, K.; Zong, L.; Tong, S.; Wang, X.; Feng, G. Gold supported on ceria nanotubes for CO oxidation. Appl. Surf. Sci. 2017, 416, 183–190. [Google Scholar] [CrossRef]
  24. Acosta, B.; Smolentseva, E.; Beloshapkin, S.; Rangel, R.; Estrada, M.; Fuentes, S.; Simakov, A. Gold supported on ceria nanoparticles and nanotubes. Appl. Catal. A Gen. 2012, 449, 96–104. [Google Scholar] [CrossRef]
  25. Fonseca, J.; Royer, S.; Bion, N.; Pirault-Roy, L.; do Carmo Rangel, M.; Duprez, D.; Epron, F. Preferential CO oxidation over nanosized gold catalysts supported on ceria and amorphous ceria-alumina. Appl. Catal. B Environ. 2012, 128, 10–20. [Google Scholar] [CrossRef]
  26. Aneggi, E.; Boaro, M.; Colussi, S.; de Leitenburg, C.; Trovarelli, A. Ceria-based materials in catalysis: Historical perspective and future trends. In Handbook of the Physics and Chemistry of Rare Earths; Bünzli, J.C., Pecharsky, V., Elsevier, B.V., Eds.; Elsevier: Amsterdam, The Netherlands, 2016; Volume 50, pp. 209–242. [Google Scholar] [CrossRef]
  27. Ansari, S.A.; Khan, M.M.; Ansari, M.O.; Kalathil, S.; Lee, J.; Cho, M.H. Band-gap engineering of CeO2 nanostructure using an electrochemically active biofilm for visible light applications. RSC Adv. 2014, 4, 16782–16791. [Google Scholar] [CrossRef]
  28. Wu, T.S.; Chen, Y.W.; Weng, S.C.; Lin, C.N.; Lai, C.H.; Huang, Y.J.; Jeng, H.T.; Chang, S.L.; Soo, Y.L. Dramatic band gap reduction incurred by dopant coordination rearrangement in Co-doped nanocrystals of CeO2. Sci. Rep. 2017, 7, 2–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Spezzati, G.; Benavidez, A.D.; DeLaRiva, A.T.; Su, Y.; Hofmann, J.P.; Asahina, S.; Olivier, E.J.; Neethling, J.H.; Miller, J.T.; Datye, A.K.; et al. CO oxidation by Pd supported on CeO2 (100) and CeO2 (111) facets. Appl. Catal. B Environ. 2019, 243, 36–46. [Google Scholar] [CrossRef]
  30. Li, J.; Liu, X.; Sun, Z.; Sun, Y.; Pan, L. Novel yolk-shell structure bismuth rich bismuth molybdate microsphere for enhanced visible-light photocatalysis. J. Colloid Interface Sci. 2015, 452, 109–115. [Google Scholar] [CrossRef] [PubMed]
  31. Panagiotopoulou, P.; Verykios, X.E. Mechanistic aspects of the low temperature steam reforming of ethanol over supported Pt catalysts. Int. J. Hydrogen Energy 2012, 37, 16333–16345. [Google Scholar] [CrossRef]
  32. Ferrer, V.; Finol, D.; Ramos, M. Caracterización química y actividad catalítica de óxidos mixtos soportados basados en cerio. Ciencia 2010, 18, 209–219. [Google Scholar]
  33. Ratova, M.; Kelly, P.J.; West, G.T.; Xia, X.; Gao, Y. Deposition of visible light active photocatalytic bismuth molybdate thin films by reactive magnetron sputtering. Materials 2016, 9, 67–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Ayame, A.; Uchida, K.; Iwataya, M.; Miyamoto, M. X-ray photoelectron spectroscopic study on α-and γ-bismuth molybdate surfaces exposed to hydrogen, propene, and oxygen. Appl. Catal. A Gen. 2002, 227, 7–17. [Google Scholar] [CrossRef]
  35. Mun, C.; Ehrhardt, J.J.; Lambert, J.; Madic, C. XPS investigations of ruthenium deposited onto representative inner surfaces of nuclear reactor containment buildings. Appl. Surf. Sci. 2007, 253, 7613–7621. [Google Scholar] [CrossRef] [Green Version]
  36. Yeung, H.; Chan, H.; Takoudis, C.G.; Weaver, M.J. High-pressure oxidation of ruthenium as probed by surface-enhanced raman and x-ray photoelectron spectroscopies. J. Catal. 1997, 172, 336–345. [Google Scholar] [CrossRef]
  37. Prusty, D.; Pathak, A.; Mukherjee, M.; Mukherjee, B.; Chowdhury, A. TEM and XPS studies on the faceted nanocrystals of Ce0.8Zr0.2O2. Mater. Charact. 2015, 100, 31–35. [Google Scholar] [CrossRef]
  38. Aguilar-Guerrero, V.; Lobo-Lapidus, R.J.; Gates, B.C. Genesis of cerium oxide supported gold catalyst for co oxidation: Transformation of mononuclear gold complexes into clusters as characterized by x-ray absorption spectroscopy. J. Phys. Chem. C 2009, 113, 3259–3269. [Google Scholar] [CrossRef]
  39. Kosacki, I.; Petrovsky, V.; Anderson, H.U. Raman spectroscopy of nanocrystalline ceria and zirconia thin films. J. Am. Ceram. Soc. 2002, 85, 2646–2650. [Google Scholar] [CrossRef]
  40. Bulushev, D.; Yuranov, I.; Suvorova, E.; Buffat, P.; Kiwi-Minsker, L. Highly dispersed gold on activated carbon fibers for low-temperature CO oxidation. J. Catal. 2004, 224, 8–17. [Google Scholar] [CrossRef] [Green Version]
  41. Santos, V.P.; Carabineiro, S.A.C.; Bakker, J.J.W.; Soares, O.S.G.P.; Chen, X.; Pereira, M.F.R.; Órfão, J.J.M.; Figueiredo, J.L.; Gascon, J.; Kapteijn, F. Stabilized gold on cerium-modified cryptomelane: Highly active in low-temperature CO oxidation. J. Catal. 2014, 309, 58–65. [Google Scholar] [CrossRef]
  42. Adhikari, R.; Joshie, B.; García, R.N.; de la Rosa, E.; Lee, W.S. Microwave hydrothermal synthesis and infrared to visible upconversion luminescence of Er3+/Yb3+ co-doped bismuth molybdate nanopowder. J. Lumin. 2016, 145, 866–871. [Google Scholar] [CrossRef]
  43. Sleight, A.; Chen, H. Structure of Bi2Mo2O9: A selective oxidation catalyst. J. Solid State Chem. 1986, 63, 70–75. [Google Scholar]
  44. Casaletto, M.P.; Longo, A.; Martorana, A.; Prestianni, A.; Venezia, A.M. XPS study of supported gold catalysts: The role of Au0 and Au+g species as active sites. Surf. Interface Anal. 2006, 38, 215–218. [Google Scholar] [CrossRef]
  45. Piotrowski, T.; Accinno, D.J. Metallography of the precious metals. Metallography 1977, 10, 243–289. [Google Scholar] [CrossRef]
  46. Carabineiro, S.A.C.; Silva, A.M.T.; Draić, G.; Tavares, P.B.; Figueiredo, J.L. Gold nanoparticles on ceria supports for the oxidation of carbon monoxide. Catal. Today 2010, 154, 21–30. [Google Scholar] [CrossRef]
  47. Amano, F.; Nogami, K.; Abe, R.; Ohtani, B. Preparation and characterization of bismuth tungstate polycrystalline flake-ball particles for photocatalytic reactions. J. Phys. Chem. C 2008, 112, 9320–9326. [Google Scholar] [CrossRef]
  48. Liao, P.C.; Mar, S.Y.; Ho, W.S.; Huang, Y.S.; Tiong, K.K. Characterization of RuO2 thin films deposited on Si by metal-organic chemical vapor deposition. Thin Solid Film. 1996, 287, 74–79. [Google Scholar] [CrossRef]
  49. Mukherjee, D.; Rao, B.G.; Reddy, B.M. CO and soot oxidation activity of doped ceria: Influence of dopants. Appl. Catal. B Environ. 2016, 197, 105–115. [Google Scholar] [CrossRef]
  50. Qadir, K.; Joo, S.H.; Mun, B.S.; Butcher, D.R.; Renzas, J.R.; Aksoy, F.; Liu, Z.; Somorjai, G.A.; Park, J.Y. Intrinsic relation between catalytic activity of CO oxidation on Ru nanoparticles and Ru oxides uncovered with ambient pressure XPS. Nano Lett. 2012, 12, 5761–5768. [Google Scholar] [CrossRef] [PubMed]
  51. Venkataswamy, P.; Jampaiah, D.; Mukherjee, D.; Aniz, C.U.; Reddy, B.M. Mn-doped ceria solid solutions for CO oxidation at lower temperatures. Catal. Lett. 2016, 146, 2105–2118. [Google Scholar] [CrossRef]
  52. Mandapaka, R.; Madras, G. Zinc and platinum co-doped ceria for WGS and CO oxidation. Appl. Catal. B Environ. 2017, 211, 137–147. [Google Scholar] [CrossRef]
  53. Zanella, R. Metodologías para la síntesis de nanopartículas: Controlando forma y tamaño, Mundo Nano. Rev. Interdiscip. Nanocienc. Nanotecnol. 2012, 5, 69–81. [Google Scholar]
  54. Duan, Y.; Li, Z.; Li, Y.; Zhang, Y.; Li, L.; Li, J. New insight of the Mars-van Krevelen mechanism of the CO oxidation by gold catalyst on the ZnO (101) surface. Comput. Theor. Chem. 2017, 1100, 28–33. [Google Scholar] [CrossRef]
  55. Zhang, S.; Li, X.; Chen, B.; Zhu, X.; Shi, C.; Zhu, A. CO oxidation activity at room temperature over Au/CeO2 catalysts: Disclosure of induction period and humidity effect. ACS Catal. 2014, 4, 3481–3489. [Google Scholar] [CrossRef]
  56. Kimling, J.; Maier, M.; Okenve, B.; Kotaidis, V.; Ballot, H.; Plech, A. Turkevich method for gold nanoparticle synthesis revisited. J. Phys. Chem. B 2006, 110, 15700–15707. [Google Scholar] [CrossRef] [PubMed]
  57. Bueno-López, A.; Krishna, K.; Makkee, M.; Moulijn, J.A. Active oxygen from CeO2 and its role in catalyzed soot oxidation. Catal. Lett. 2005, 99, 203–205. [Google Scholar] [CrossRef]
Figure 1. Scanning electron microscopy (SEM) images of (a) Bi2Mo0.95Ru0.05O6 (MBR1); (b) Bi2Mo0.97Ru0.03O6 (MBR2); (c) CeO2/Bi2Mo0.95Ru0.05O6 (OCMBR1); (d) CeO2/Bi2Mo0.97Ru0.03O6 (OCMBR2); (e) Au/Bi2Mo0.95Ru0.05O6 (AuMBR1); and (f) Au/Bi2Mo0.97Ru0.03O6 (AuMBR2) compounds under a magnification of 20,000X.
Figure 1. Scanning electron microscopy (SEM) images of (a) Bi2Mo0.95Ru0.05O6 (MBR1); (b) Bi2Mo0.97Ru0.03O6 (MBR2); (c) CeO2/Bi2Mo0.95Ru0.05O6 (OCMBR1); (d) CeO2/Bi2Mo0.97Ru0.03O6 (OCMBR2); (e) Au/Bi2Mo0.95Ru0.05O6 (AuMBR1); and (f) Au/Bi2Mo0.97Ru0.03O6 (AuMBR2) compounds under a magnification of 20,000X.
Catalysts 09 00947 g001
Figure 2. X-ray diffraction patterns of (a) AuMBR1; (b) AuMBR2; (c) OCMBR2; (d) OCMBR1; (e) MBR2; and (f) MBR1 synthesized catalysts. Cu Kα radiation operating with a step size of 0.02°/min from 10° to 90°.
Figure 2. X-ray diffraction patterns of (a) AuMBR1; (b) AuMBR2; (c) OCMBR2; (d) OCMBR1; (e) MBR2; and (f) MBR1 synthesized catalysts. Cu Kα radiation operating with a step size of 0.02°/min from 10° to 90°.
Catalysts 09 00947 g002
Figure 3. General X-ray photoelectron spectroscopy (XPS) scan of (a) MBR1 and (b) MBR2 support catalysts.
Figure 3. General X-ray photoelectron spectroscopy (XPS) scan of (a) MBR1 and (b) MBR2 support catalysts.
Catalysts 09 00947 g003
Figure 4. High-resolution XPS analysis for Ru3d of (a) MBR1 and (b) MBR2 supports.
Figure 4. High-resolution XPS analysis for Ru3d of (a) MBR1 and (b) MBR2 supports.
Catalysts 09 00947 g004
Figure 5. General XPS scans of (a) OCMBR1 and (b) OCMBR2 catalysts.
Figure 5. General XPS scans of (a) OCMBR1 and (b) OCMBR2 catalysts.
Catalysts 09 00947 g005
Figure 6. General XPS scans of (a) AuMBR1 and (b) AuMBR2 catalysts.
Figure 6. General XPS scans of (a) AuMBR1 and (b) AuMBR2 catalysts.
Catalysts 09 00947 g006
Figure 7. High-resolution XPS analysis of Au4f of (a) AuMBR1; (b) AuMBR1_R; (c) AuMBR2; and (d) AuMBR2_R (R: reduced).
Figure 7. High-resolution XPS analysis of Au4f of (a) AuMBR1; (b) AuMBR1_R; (c) AuMBR2; and (d) AuMBR2_R (R: reduced).
Catalysts 09 00947 g007
Figure 8. Raman spectra of (a) MBR2; (b) MBR1; and (c) Bi2MoO6 compounds.
Figure 8. Raman spectra of (a) MBR2; (b) MBR1; and (c) Bi2MoO6 compounds.
Catalysts 09 00947 g008
Figure 9. Comparative graph of CO conversion percentage for MB, MBR1, and MBR2 catalysts.
Figure 9. Comparative graph of CO conversion percentage for MB, MBR1, and MBR2 catalysts.
Catalysts 09 00947 g009
Figure 10. Comparative graph of CO conversion for OCMBR1, OCMBR2, AuMBR1_R, AuMBR2_R, AuMBR1, and AuMBR2 catalysts. (R: reduced).
Figure 10. Comparative graph of CO conversion for OCMBR1, OCMBR2, AuMBR1_R, AuMBR2_R, AuMBR1, and AuMBR2 catalysts. (R: reduced).
Catalysts 09 00947 g010
Figure 11. Schematic flow diagram of the reaction system.
Figure 11. Schematic flow diagram of the reaction system.
Catalysts 09 00947 g011
Table 1. Elemental chemical analysis (atomic percent) obtained by energy-dispersive X-ray spectroscopy (EDS) using an EDAX detector.
Table 1. Elemental chemical analysis (atomic percent) obtained by energy-dispersive X-ray spectroscopy (EDS) using an EDAX detector.
SampleChemical Composition (Atomic Percent)
BiMoRuOCeAu
MBR122.0513.621.5664.34--
MBR229.8112.561.4957.63--
OCMBR116.113.652.1264.673.46-
OCMBR216.7213.471.8964.953.21-
AuMBR116.0419.341.6961.73-1.2
AuMBR214.4922.291.3560.54-1.33
Table 2. Quantification of chemical gold species before and after reduction with H2.
Table 2. Quantification of chemical gold species before and after reduction with H2.
CatalystsAu Species (%)
Au+3Au0
AuMBR1Before reduction61.3438.66
After reduction39.2460.76
AuMBR2Before reduction45.0354.97
After reduction38.5661.44

Share and Cite

MDPI and ACS Style

González, E.E.; Rangel, R.; Lara, J.; Bartolo-Pérez, P.; Alvarado-Gil, J.J.; Galván, D.H.; García, R. {CeO2/Bi2Mo1−xRuxO6} and {Au/Bi2Mo1−xRuxO6} Catalysts for Low-Temperature CO Oxidation. Catalysts 2019, 9, 947. https://doi.org/10.3390/catal9110947

AMA Style

González EE, Rangel R, Lara J, Bartolo-Pérez P, Alvarado-Gil JJ, Galván DH, García R. {CeO2/Bi2Mo1−xRuxO6} and {Au/Bi2Mo1−xRuxO6} Catalysts for Low-Temperature CO Oxidation. Catalysts. 2019; 9(11):947. https://doi.org/10.3390/catal9110947

Chicago/Turabian Style

González, Edson Edain, Ricardo Rangel, Javier Lara, Pascual Bartolo-Pérez, Juan José Alvarado-Gil, Donald Homero Galván, and Rafael García. 2019. "{CeO2/Bi2Mo1−xRuxO6} and {Au/Bi2Mo1−xRuxO6} Catalysts for Low-Temperature CO Oxidation" Catalysts 9, no. 11: 947. https://doi.org/10.3390/catal9110947

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