Next Article in Journal
Application of Various Metal-Organic Frameworks (MOFs) as Catalysts for Air and Water Pollution Environmental Remediation
Next Article in Special Issue
Ceria-Based Catalysts Studied by Near Ambient Pressure X-ray Photoelectron Spectroscopy: A Review
Previous Article in Journal
Friedel–Crafts-Type Alkylation of Indoles in Water Using Amphiphilic Resin-Supported 1,10-Phenanthroline–Palladium Complex under Aerobic Conditions
Previous Article in Special Issue
In Situ Investigations on the Facile Synthesis and Catalytic Performance of CeO2-Pt/Al2O3 Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Evoked Methane Photocatalytic Conversion to C2 Oxygenates over Ceria with Oxygen Vacancy

1
College of Sciences, Shanghai University, Shanghai 200444, China
2
CAS Key Laboratory of Low-carbon Conversion Science and Engineering, Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai 201210, China
3
School of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(2), 196; https://doi.org/10.3390/catal10020196
Submission received: 27 December 2019 / Revised: 17 January 2020 / Accepted: 31 January 2020 / Published: 6 February 2020

Abstract

:
Direct conversion of methane to its oxygenate derivatives remains highly attractive while challenging owing to the intrinsic chemical inertness of CH4. Photocatalysis arises as a promising green strategy which could stimulate water splitting to produce oxidative radicals for methane C–H activation and subsequent C–C coupling. However, synthesis of a photocatalyst with an appropriate capability of methane oxidation by water remains a challenge using an effective and viable approach. Herein, ceria nanoparticles with abundant oxygen vacancies prepared by calcinating commercial CeO2 powder at high temperatures in argon are reported to capably produce ethanol and aldehyde from CH4 photocatalytic oxidation under ambient conditions. Although high-temperature calcinations lead to lower light adsorptions and increased band gaps to some extent, deficient CeO2 nanoparticles with oxygen vacancies and surface CeIII species are formed, which are crucial for methane photocatalytic conversion. The ceria catalyst as-calcinated at 1100 °C had the highest oxygen vacancy concentration and CeIII content, achieving an ethanol production rate of 11.4 µmol·gcat−1·h−1 with a selectivity of 91.5%. Additional experimental results suggested that the product aldehyde was from the oxidation of ethanol during the photocatalytic conversion of CH4.

1. Introduction

Methane, as the main constituent of natural gas and shale gas, is widely used as an important and clean fuel [1]. With large-scale exploitation of unconventional natural gas, especially shale gas, direct conversion of methane as a chemical feedstock to valuable chemicals is highly attractive [2,3]. Nevertheless, the CH4 molecule exhibits a perfectly symmetrical tetrahedral structure with negligible electron affinity and the highest C–H bond strength (434 ± 6 kJ·mol−1) in all hydrocarbons, implying great difficulties in activating methane and breaking C–H bonds [4,5,6,7]. To date, large-scale methane conversion applications are limited in thermal catalysis under harsh reaction conditions. For instance, methane reformation occurs at high temperature with complicated facilities [8,9], where methane is converted only to syngas (CO/H2) as a feedstock for the Fischer–Tropsch synthesis process [10,11]. Although oxidative coupling of methane can produce C2 hydrocarbons such as ethylene and ethane [12,13], the existence of oxygen at high temperatures may present potential explosion risks and irreversible overoxidation to generate CO2 with low carbon utilization [14]. The highly promising non-oxidative coupling of methane to olefins and aromatics recently reported [15,16] requires high temperatures (~1000 °C) to activate methane. It is highly desired to develop a new approach for methane conversion.
Using renewable solar energy to drive methane conversion by photocatalytic processes remains highly attractive, as they feature both solar energy storage and direct utilization of methane as feedstock. Moreover, water is the most naturally abundant source of oxidative species, which is available and low-cost for CH4 photooxidation [17]. Photocatalysis arises as a promising green strategy which could stimulate water splitting to produce oxidative radicals for methane C–H activation and subsequent C–C coupling [17,18,19,20]. Synthesis of photocatalysts with appropriate capabilities of water splitting and methane oxidation remains a challenge under an effective and viable approach.
In the past few decades, cerium-based materials gained increasing attention as photocatalysts alternative to conventional titania, both for wastewater treatment [21,22,23,24,25] and water splitting [26,27,28]. As a representative cerium-based material, ceria is a wide-band-gap semiconductor (3.2–3.4 eV), which is used in various industrial applications such as solar energy for water splitting [29], as an oxygen storage material [30], and in solar cells, luminescent materials, and photocatalytic applications [31,32,33]. The usability of ceria for most of these applications depends on its ability to release or take up oxygen. Ceria has large oxygen storage capacity depending on Ce3+/Ce4+ redox cycles. More recently, some studies [34,35] reported that increasing the surface defect concentration of ceria by calcination enhanced the photocatalytic degradation of dyes. The surface defects such as oxygen vacancies can promote the generation of hydroxyl radicals from water splitting under light irradiation. Based on this principle, modification of oxygen vacancy defects on ceria is theoretically beneficial for the photocatalytic oxidation of methane with water. Herein, we report commercially available ceria nanoparticles that behave inertly for methane photocatalytic conversion, which are then evoked and become active via high-temperature calcination in Ar atmosphere. Oxygen vacancies and surface CeIII species in the as-prepared ceria catalysts are proposed to be crucial for generating oxidative radicals for methane oxidation under simulated solar light irradiation. The present study illustrates a simple and effective approach for fabricating active ceria catalysts for direct methane conversion and solar energy utilization fields.

2. Results and Discussion

2.1. Characterizations of the Catalysts

A series of ceria catalysts were synthesized by simple calcination of commercial CeO2 nanoparticles in Ar at 500–1100 °C (see Section 3.1 for details). The oxygen signal (m/z = 32) from a temperature-programmed calcination indicated that lattice oxygen removal started around 600 °C (Figure S1A, Supplementary Materials), and the actual variation of the CeO2 crystal structure potentially occurred at lower temperature. The thermogravimetry (TG) curve of CeO2-raw (Figure S1B, Supplementary Materials) showed a very slow mass decrease until 580 °C, followed by rapid mass loss in the range of 580–1040 °C, implying lattice oxygen removal in Ar flow. The TG results further indicate that an inert atmosphere is more beneficial for oxygen removal to form oxygen vacancies, compared to the oxygen-rich atmosphere (Figure S1B, Supplementary Materials). At 500 °C and higher temperatures with an interval of 100 °C, CeO2-x catalysts were obtained, where x represents the preserved calcination temperature. Figure 1 shows the typical photographs of as-prepared CeO2-x samples at different temperatures. The yellowish colors of CeO2-500, CeO2-600, and CeO2-700 seem highly similar to that of CeO2-raw. Apparent color darkening started from CeO2-800, and then gradually changed to brown with increasing calcination temperature. A possible reason may be derived from the variation of lattice structures, especially oxygen vacancy (vide infra), in agreement with previous reports [34].
X-ray diffraction (XRD) patterns of CeO2-raw and CeO2-x are displayed in Figure 2. The diffraction peaks corresponding to (111), (200), (220), (311), (222), (400), (331), and (420) planes of cubic ceria with a space group of Fm-3m (JCPDS no.34-0394) were detected, and no peaks from the other ceric and cerous oxide species were manifested in all samples. This implies that the high-temperature calcinations, even as high as 1100 °C, in Ar did not cause phase transformation. The average particle sizes of CeO2-x were affected by the calcination temperature, especially high temperatures. The particle size of CeO2-500 (18.4 nm), as calculated using the Scherrer equation on the basis of the (111) peak, was close to that of CeO2-raw (18.0 nm), although CeO2-500 was subjected to the calcination treatment of 500 °C. Subsequently, the particle sizes of CeO2-x exponentially increased with increasing calcination temperature, that is, particle sizes from 20.7 nm of CeO2-600 rapidly increased to 29.3 nm (CeO2-700), 48.5 nm (CeO2-800), 65.9 nm (CeO2-900), 78.7 nm (CeO2-1000), and 93.2 nm (CeO2-1100).
According to previous reports [35,36,37,38], the total concentrations of oxygen vacancy in CeO2-raw and CeO2-x can be calculated using Equations (6) and (7) (vide infra, see Section 3 for full details). With increasing calcination temperatures, the concentrations of oxygen vacancy of CeO2-x slowly increase initially, and then rapidly increase at higher temperatures, as shown in Figure S2 (Supplementary Materials). The oxygen vacancy concentration of CeO2-500 was 0.6 × 1018 vacancies/cm3, only slightly higher than that of CeO2-raw (0.3 × 1018 vacancies/cm3). CeO2-600 had a much larger oxygen vacancy concentration (2.1 × 1018 vacancies/cm3). From the catalyst CeO2-700 (6.5 × 1018 vacancies/cm3), the oxygen vacancy concentrations of CeO2-x further linearly increased to 13.8 × 1018 vacancies/cm3 (CeO2-800), 19.6 × 1018 vacancies/cm3 (CeO2-900), 26.2 × 1018 vacancies/cm3 (CeO2-1000), and 32.5 × 1018 vacancies/cm3 (CeO2-1100). These results demonstrate that the high-temperature calcinations in Ar facilitated the formation of oxygen vacancies. The high concentrations of oxygen vacancies from the calcination at 800 °C also led to a darkened yellowish appearance, as recognized previously [39].
The surface electronic states of CeO2-raw and CeO2-x were studied by X-ray photoelectron spectroscopy (XPS), as shown in Figure 3A. The Ce 3d spectrum of CeO2-raw shows the main Ce 3d5/2 and 3d3/2 core peaks at binding energies of 881.6 and 901.8 eV, respectively, indicating the CeIV state in CeO2. Two CeO2 shake-up satellite 3d5/2 peaks including the CeIV satellite (i) and CeIV satellite (ii) at 889.7 and 897.5 eV, respectively, also appeared for CeO2-raw. As for CeO2-x, in addition to CeO2 peaks, new peaks located at 879.6 (Ce 3d5/2) and 897.7 eV (Ce 3d3/2) were the characteristic CeIII peaks, along with the CeIII satellite 3d5/2 peak at 886.0 eV. However, XRD characterizations showed only CeO2 phase as the bulk composition in all CeO2-x catalysts, implying that the surface CeIII oxide layers evidenced by XPS were too thin to be detectable. Variations of the relative intensity ratios of CeIII/CeIV with respect to CeO2-raw and individual CeO2-x are displayed in Figure 3B. In the lower calcination temperature range from 500 °C to 800 °C, the CeIII/CeIV ratios increased relatively quickly from 0.08 (CeO2-500) to 0.33 (CeO2-800). Subsequently, the CeIII/CeIV ratios continuously increased to 0.5 (CeO2-1100) with a relatively slow rate, implying potential surface/interface configuration inflexion in CeO2-800.
The absorption edges and band properties for CeO2-raw and CeO2-x were studied using UV–Vis diffuse reflectance spectra (DRS) (Figure 4). One can see that CeO2-raw showed an intense adsorption band from 200 nm to 380 nm and a trailing adsorption band until 480 nm. As for CeO2-x, the intensities of the low-frequency region gradually decreased with increasing calcination temperatures, while the trailing adsorption widths of CeO2-x in the high-frequency region fluctuated to some extent, implying the effective influences of calcination on the band structures of ceria. Tauc plots based on DRS absorption edges show that the band gaps were sharply broadened from 2.97 eV of CeO2-raw to 3.01 eV (CeO2-500), 3.03 eV (CeO2-600), 3.05 eV (CeO2-700), 3.08 eV (CeO2-800), 3.12 eV (CeO2-900), 3.15 eV (CeO2-1000), and 3.18 eV (CeO2-1000) (inset I in Figure 4). The band-gap values increased almost linearly with increasing calcination temperatures (inset II in Figure 4). The oxygen vacancies derived from oxygen atom effusion during the calcination process caused the band gaps to broaden stepwise [40,41], implying the effective influence of calcination on the band-gap structures of ceria.
All CeO2-raw and CeO2-x catalysts exhibited excellent photocurrent response abilities in the light on and off cycles, as demonstrated in Figure 5. Moreover, the photocurrent density of CeO2-raw was about 0.061 µA·cm−2 in the first light on stage, before gradually decreasing to 0.041 µA·cm−2 in the last light on stage. The photocurrent densities of CeO2-x were larger than those of CeO2-raw in all light on stages. Furthermore, the intensities of the photocurrent densities of CeO2-x stepwise increased with increasing calcination temperatures. The initial photocurrent densities of CeO2-x were 0.085 µA·cm−2 (CeO2-500), 0.10 µA·cm−2 (CeO2-600), 0.115 µA·cm−2 (CeO2-700), 0.120 µA·cm−2 (CeO2-800), 0.137 µA·cm−2 (CeO2-900), 0.163 µA·cm−2 (CeO2-1000), and 0.183 µA·cm−2 (CeO2-1100). CeO2-x also showed attenuating photocurrent densities in the following light on stages, and CeO2-800 retained a relatively slow attenuating rate compared to the other CeO2-x catalysts. These results adequately prove that calcination can effectively affect the band structures of ceria, resulting in the promoted photo-response.

2.2. Photocatalytic Conversion of CH4

Photocatalytic methane oxidation reactions were conducted over CeO2-raw and CeO2-x catalysts stirred/dispersed in pure water under simulated sunlight irradiation. As shown in Figure 6, the aldehyde and ethanol products were reproducibly generated, and no other products were detected over CeO2-x catalysts. In contrast, no products were generated over CeO2-raw from CH4 photooxidation. CeO2-x produced a small amount of aldehyde with relatively stable production rates of about 1.0 µmol·gcat−1·h−1 with ~10% fluctuation. Note that ethanol production rates over CeO2-x catalysts quickly increased from 0.9 µmol·gcat−1·h−1 (CeO2-500) to 2.3 µmol·gcat−1·h−1 (CeO2-600), 4.6 µmol·gcat−1·h−1 (CeO2-700), and 8.5 µmol·gcat−1·h−1 (CeO2-800) initially, and then further increased to 9.4 µmol·gcat−1·h−1 (CeO2-900), 10.3 µmol·gcat−1·h−1 (CeO2-1000), and 11.4 µmol·gcat−1·h−1 (CeO2-1100) with relatively slow rates at the higher temperature range. Such variations of aldehyde and ethanol production rates over CeO2-x resulted in an asymptotic curve for ethanol selectivity. That is, the ethanol selectivity increased quickly from 50.2% (CeO2-500) to 71.3% (CeO2-600), 80.9% (CeO2-700), and 88.0% (CeO2-800), and then slowly to 90.4% (CeO2-900), 90.7% (CeO2-1000), 91.5% (CeO2-1100). Note that the changes in the ethanol selectivity and production rate were highly similar to those of CeIII/CeIV ratios in CeO2-x, indicating that the activity of producing ethanol was sensitive to the surface CeIII species on the as-prepared ceria. Recently, Murali et al. reported that an optimal CeOx layer in the ceria nanoparticle catalyst was obtained via tuning the molar ratio of CeIII, exhibiting an enhanced photocatalytic performance [34,35,37,42]. The quantum efficiency of ethanol production over the catalyst CeO2-1100 was about 0.3% (Table S1, Supplementary Materials).
Based on the results and discussion above, we propose a hypothetic radical mechanism for photocatalytic CH4 conversion over CeO2-x catalysts according to previous reports [17,34]. Oxidative species O2− and/or hydroxyl radicals (∙OH) are firstly formed on the surface of CeO2-x catalysts, which are derived from the interaction between water and holes induced by active oxygen vacancies under simulated solar light irradiation. These ∙OH radicals interact with CH4 to produce methyl radicals (∙CH3) (Equation (1)), and then some of the as-formed ∙CH3 undergoes radical coupling to generate ethane (Equation (2)), which is further attacked by ∙OH to produce ethyl radicals (∙CH2CH3) (Equation (3)). Ethanol is produced from the reaction between ∙CH2CH3 and H2O (Equation (4)), while the reaction between hydrogen radicals (∙H) and ∙OH to produce water (Equation (5)) is a chain termination reaction.
C H 4   +   · O H     · C H 3   +   H 2 O .  
· C H 3   +   · C H 3     C 2 H 6 .  
C 2 H 6   +   · O H     · C H 2 C H 3   +   H 2 O .
· C H 2 C H 3   +   H 2 O     C 2 H 5 O H   +   · H .  
· H   +   · O H     H 2 O .
Furthermore, it was demonstrated that the formation of aldehyde is attributed to the oxidation of the product ethanol under the photocatalytic conditions via additional experiments of photocatalytic ethanol conversion. As shown in Figure S3 (Supplementary Materials), the initial ethanol concentrations were 1.03, 1.54, 2.57, and 3.86 µmol∙L−1 in Test 1, Test 2, Test 3, and Test 4, respectively. After simulated solar light irradiation for 2 h with Ar flow instead of methane flow, the post-reaction liquid products were found to be aldehyde and unreacted ethanol. Moreover, the amount of aldehyde in different tests remained almost constant (0.23 µmol∙L−1). Note that the sums of the unreacted ethanol and as-generated aldehyde were equal to the initial ethanol in the individual test. This implies that certain amount of the product ethanol was converted to the equivalent aldehyde under the CH4 photocatalytic conditions. That is, the oxidative species also led to the occurrence of an ethanol oxidation side reaction, resulting in a nearly constant amount of aldehyde in the CH4 photocatalytic conversion process. The cycling performance tests showed that the production rates of aldehyde over CeO2-1100 remained constant (maintaining at 1.1 µmol∙gcat−1∙h−1) at different cycles (Figure S4, Supplementary Materials). The ethanol production rates were relatively stable with only about 6% stepwise decay from Cycle 1 to Cycle 3, but they decreased rapidly at Cycle 4 with about 20% decay. In addition, the catalyst CeO2-1100 showed comparable photocatalytic activities with the best photocatalysts under mild reaction conditions (Table S2, Supplementary Materials).

3. Materials and Methods

3.1. Preparation of Catalysts

In this work, a calcination method in Ar atmosphere was adopted, and the details were as follows: commercial CeO2 samples (about 0.1 g) were put in a U-type quartz tube and purged with Ar (99.999%, 200 mL∙min1) for 30 min before heating. Then, the samples were calcined in flowing Ar at a heating rate of 5 °C∙min1 to target temperatures (500 °C, 600 °C, 700 °C, 800 °C, 900 °C, 1000 °C, and 1100 °C) and held at the individual target temperatures for 4 h. After cooling down to room temperature in Ar flow, the as-obtained catalysts were stored in airtight glass bottles.

3.2. Physical Characterizations

The crystal structures of the catalysts were determined by X-ray diffraction (XRD) patterns performed on an X-ray diffractometer (Ultima IV, Rigaku Corporation, Tokyo, Japan) employing Cu Kα radiation (λ = 1.54056 Å) at 40 kV and 40 mA. XRD patterns were recorded at a scan rate of 5°∙min−1 and in the range of 20°–80°. The concentrations of oxygen vacancy were calculated using Equations (6) and (7).
3 4 ( a a 0 ) = c [ r C e 3 + r C e 4 + + 1 4 ( r V 0 r O 2 ) ] ,
V 0 = 2 c ,
where a is the lattice constant based on the full width half maximum (FWHM) of the (111) diffraction peak of individual samples, and the a values for CeO2-x were 0.5411 nm (CeO2-raw), 0.54112 nm (CeO2-500), 0.54115 nm (CeO2-600), 0.54126 nm (CeO2-700), 0.54145 nm (CeO2-800), 0.5416 nm (CeO2-900), 0.54177 nm (CeO2-1000), and 0.54192 nm (CeO2-1100); a 0 is the lattice parameter (0.5411 nm) of CeO2 (JCPDS no.34-0394), c is the coefficient of vacancy concentration, r C e 3 + and r C e 4 + are the radii of Ce3+ and Ce4+, respectively (   r C e 3 + = 0.1283 nm,   r C e 4 + = 0.1098 nm), r V 0 and r O 2 are the radii of oxygen vacancy and O2−, respectively ( r V 0 = 0.138 nm, r O 2 = 0.124 nm), and   V 0 is the oxygen vacancy concentration. UV–visible absorption spectra were recorded on a UV–visible spectrophotometer (UV-2700, Shimadzu Co., Ltd., Kyoto, Japan). XPS spectra were obtained using an X-ray photoelectron spectrometer (Quantum 2000 Scanning ESCA Microprobe instrument, Thermo Fisher Scientific Inc., Waltham, MA., USA) with a monochromatic excitation source of Al Kα radiation ( = 1486.6 eV) performed under 12 kV and 4 mA. The binding energies in all XPS spectra were calibrated according to the C 1s peak (284.8 eV). The XPS spectra were deconvolved using a commercially available data-fitting program (Advantage 5.967 software (2016), Thermo Fisher Scientific Inc., Waltham, MA, USA) after a Shirley background subtraction procedure. The photocurrent response tests of the catalysts were carried out on an electrochemical workstation (VMP3, Bio-Logic Inc., Seyssinet-Pariset, France) using the chronoamperometry method at a fixed potential of 0 mV vs. Ag/AgCl under chopped illumination periods of 60 s. The catalysts coated on the carbon paper pieces (2 cm × 2 cm) were used as the working electrode, a Pt wire was used as the counter-electrode, and KCl-saturated Ag/AgCl was used as the reference electrode. The CH4-saturated 0.1 M NaOH aqueous solution was used as the electrolyte solution.

3.3. Photocatalytic Activity Tests

In a typical photocatalytic CH4 oxidation reaction, the catalyst (2 mg) was suspended in ultra-pure water (15 mL) with continuous magnetic stirring in a gas-tight glass cell equipped with a constant temperature circulator (25 °C). The set-up of the CH4 photocatalytic conversion is shown in Figure S5 (Supplementary Materials), and the glass cell with a volume of 20 mL was covered with a quartz window (effective irradiation diameter of 5 cm). The pure CH4 (99.999%) was delivered into the cell at a constant rate of 4 mL·min1 and was allowed to purge for 30 min prior to the beginning of experiments. The pressure of the reaction cell remained at 0.2 bar. A 300-W Xe lamp (CEL-HXF300, Beijing CeAuLight Co., Ltd., Beijing, China) with an air mass (AM 1.5 filter, 100 mW·cm−2) was used as the simulated light source, which possessed a similar solar spectrum (Figure S6, Supplementary Materials). In the methane photocatalytic conversion, there were two kinds of products, i.e., one in the gas phase, and the other in the liquid phase. The gas-phase products were measured by the on-line GC (GC-2014, Shimadzu Co., Ltd., Kyoto, Japan) equipped with a Molecular sieve-13X 60/80 column and a Plot-Q80/100 column using the flame ionization detector (FID). The amount of the possible gas-phase products was too little to be detected by the on-line GC, resulting in no gas-phase product detected in all CH4 photocatalytic oxidation tests. On the other hand, products such as ethanol and aldehyde were dissolved in the suspension, which could not be detected by the on-line GC. Instead, these liquid-phase products were detected by the off-line GC. In brief, after filtration of the suspending particles (using an injector with the 0.02-µm polytetrafluoroethylene (PTFE) membrane), liquid products from each photocatalytic experiment for 2 h were analyzed using the off-line GC-2014 (Shimadzu) equipped with an autosampler and an OVI-G43 capillary column (Supelco®, Sigma-Aldrich Inc., St. Louis, MO, USA). The production rates of liquid products were calculated as follows:
P r o d u c t i o n   r a t e   = C p r o d u c t × V × ρ p r o d u c t M p r o d u c t × m c a t a l y s t × t ,
where C p r o d u c t (ppm) is the concentration of the product (aldehyde or ethanol), V is the liquid volume (15 mL), ρ p r o d u c t is the density of the products aldehyde (0.7834 g·cm−3) or ethanol (0.7893 g·cm−3), M p r o d u c t is the molar mass of aldehyde (44 g·mol−1) or ethanol (46.03 g·mol−1), and t is the reaction time (2 h). Control experiments including no photocatalysts or no light irradiation were carried out, and no products were detected.
Furthermore, additional experiments for investigating aldehyde origin were conducted with ethanol as an initial reactant in water under the same photocatalytic reaction conditions except for replacing methane with Ar. Briefly, the catalyst (2 mg) and specific amounts of ethanol with concentrations of 1.03, 1.54, 2.57, and 3.86 µmol·L−1 were added to ultra-pure water (15 mL) with continuous magnetic stirring in the cell. The Ar flow (4 mL·min−1) was delivered into the cell, and light irradiation was applied to the cell for 2 h after Ar purging for 30 min. The post-reaction liquid products were analyzed after filtration by the off-line GC.

4. Conclusions

In summary, we presented an effective approach for the selective conversion of methane to C2 products ethanol and aldehyde in ambient conditions and simulated solar light irradiation. Commercially available ceria nanoparticles that behave inertly for methane photocatalytic conversion were evoked and became active after modification with oxygen vacancies via calcination at high temperatures in Ar atmosphere. Higher temperatures led to quickly growing particle sizes with abundant oxygen vacancies, as well as higher ratios of CeIII/CeIV, resulting in higher activities for methane conversion. The production of ethanol was found to be more sensitive to the surface CeIII species on as-prepared ceria. The aldehyde formation can be attributed to the oxidation of the product ethanol. The present study provides an effective approach to create oxygen vacancies and surface CeIII species in commercial ceria for evoking methane photocatalytic conversion to C2 oxygenates.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/2/196/s1: Figure S1: (A) Oxygen signal (m/z = 32) of mass spectrometry in Ar flow (20 mL/min), and (B) TG curves during the temperature-programmed calcinations of commercial CeO2 nanoparticles at a heating rate of 5 °C·min−1 in air and Ar flow (20 mL/min), respectively; Figure S2: Concentrations of oxygen vacancies of CeO2-raw and CeO2-x; Figure S3: Photocatalytic ethanol conversion to aldehyde over CeO2-1100 under simulated sunlight irradiation in pure water. Initial ethanol concentrations were 1.03, 1.54, 2.57, and 3.86 µmol·L−1 in the experiments Test1, Test2, Test3, and Test4, respectively; Figure S4: Cycling performances of CH4 photocatalytic conversion over CeO2-1100; Figure S5: The photographs of (A) the methane photocatalytic conversion set-up and (B) the gas-tight glass cell; Table S1: Calculation of quantum efficiency of ethanol production over the catalyst CeO2-1100; Table S2: Comparison of the photocatalytic activities.

Author Contributions

W.C. conceptualized and designed the study. J.D. prepared catalysts and performed the reactions and most of characterizations. G.W. conducted additional tests. J.D., W.C., Y.S., X.D., G.L., J.F., W.W., and Y.S. contributed data analysis and wrote the paper. All authors read and agreed to the published version of the manuscript.

Funding

This reseach was funded by the National Natural Science Foundation of China (Nos. 91745114, 21802160), the Ministry of Science and Technology of China (Nos. 2016YFA0202800, 2018YFB0604700), the Shanghai Sailing Program (No. 18YF1425700), and the Shanghai Advanced Research Institute Innovation Research Program (Nos. Y756812ZZ1 (172002) and Y756803ZZ1 (171003)).

Acknowledgments

We appreciate the facility supports from the Shanghai Functional Platform for Innovation Low Carbon Technology. W.C. also acknowledges the support from the Hundred Talents Program of the Chinese Academy of Sciences.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yuliati, L.; Yoshida, H. Photocatalytic conversion of methane. Chem. Soc. Rev. 2008, 37, 1592–1602. [Google Scholar] [CrossRef] [PubMed]
  2. Gunsalus, N.J.; Koppaka, A.; Park, S.H.; Bischof, S.M.; Hashiguchi, B.G.; Periana, R.A. Homogeneous functionalization of methane. Chem. Rev. 2017, 117, 8521–8573. [Google Scholar] [CrossRef]
  3. Xie, S.; Lin, S.; Zhang, Q.; Tian, Z.; Wang, Y. Selective electrocatalytic conversion of methane to fuels and chemicals. J. Energy Chem. 2018, 27, 1629–1636. [Google Scholar] [CrossRef] [Green Version]
  4. Schwach, P.; Pan, X.; Bao, X. Direct conversion of methane to value-added chemicals over heterogeneous catalysts: Challenges and prospects. Chem. Rev. 2017, 117, 8497–8520. [Google Scholar] [CrossRef]
  5. Tang, P.; Zhu, Q.; Wu, Z.; Ma, D. Methane activation: The past and future. Energy Environ. Sci. 2014, 7, 2580–2591. [Google Scholar] [CrossRef]
  6. Paunović, V.; Zichittella, G.; Moser, M.; Amrute, A.P.; Pérez-Ramírez, J. Catalyst design for natural-gas upgrading through oxybromination chemistry. Nat. Chem. 2016, 8, 803–809. [Google Scholar] [CrossRef] [PubMed]
  7. Lee, K.J.; Dempsey, J.L. When electrochemistry met methane: Rapid catalyst oxidation fuels hydrocarbon functionalization. ACS Cent. Sci. 2017, 3, 1137–1139. [Google Scholar] [CrossRef] [PubMed]
  8. Nikolla, E.; Schwank, J.W.; Linic, S. Hydrocarbon steam reforming on Ni alloys at solid oxide fuel cell operating conditions. Catal. Today 2008, 136, 243–248. [Google Scholar] [CrossRef]
  9. Wang, B.; Albarracín-Suazo, S.; Pagán-Torres, Y.; Nikolla, E. Advances in methane conversion processes. Catal. Today 2017, 285, 147–158. [Google Scholar] [CrossRef] [Green Version]
  10. Enger, B.C.; Lødeng, R.; Holmen, A. A review of catalytic partial oxidation of methane to synthesis gas with emphasis on reaction mechanisms over transition metal catalysts. Appl. Catal. A Gen. 2008, 346, 1–27. [Google Scholar] [CrossRef]
  11. Olivos-Suarez, A.I.; Szécsényi, A.; Hensen, E.J.; Ruiz-Martinez, J.; Pidko, E.A.; Gascon, J. Strategies for the direct catalytic valorization of methane using heterogeneous catalysis: Challenges and opportunities. ACS Catal. 2016, 6, 2965–2981. [Google Scholar] [CrossRef]
  12. Karakaya, C.; Kee, R.J. Progress in the direct catalytic conversion of methane to fuels and chemicals. Prog. Energy Combust. Sci. 2016, 55, 60–97. [Google Scholar] [CrossRef] [Green Version]
  13. Mesters, C. A selection of recent advances in C1 chemistry. Annu. Rev. Chem. Biomol. Eng 2016, 7, 223–238. [Google Scholar] [CrossRef]
  14. Song, J.; Sun, Y.; Ba, R.; Huang, S.; Zhao, Y.; Zhang, J.; Zhu, Y. Monodisperse Sr–La2O3 hybrid nanofibers for oxidative coupling of methane to synthesize C2 hydrocarbons. Nanoscale 2015, 7, 2260–2264. [Google Scholar] [CrossRef] [PubMed]
  15. Guo, X.; Fang, G.; Li, G.; Ma, H.; Fan, H.; Yu, L.; Tan, D. Direct, nonoxidative conversion of methane to ethylene, aromatics, and hydrogen. Science 2014, 344, 616–619. [Google Scholar] [CrossRef]
  16. Soulivong, D.; Norsic, S.; Taoufik, M.; Coperet, C.; Thivolle-Cazat, J.; Chakka, S.; Basset, J.M. Non-Oxidative Coupling Reaction of Methane to Ethane and Hydrogen Catalyzed by the Silica-Supported Tantalum Hydride: (≡SiO)2Ta−H. J. Am. Chem. Soc. 2008, 130, 5044–5045. [Google Scholar] [CrossRef]
  17. Murcia-López, S.; Villa, K.; Andreu, T.; Morante, J.R. Partial oxidation of methane to methanol using bismuth-based photocatalysts. ACS Catal. 2014, 4, 3013–3019. [Google Scholar] [CrossRef]
  18. Chen, X.; Li, Y.; Pan, X.; Cortie, D.; Huang, X.; Yi, Z. Photocatalytic oxidation of methane over silver decorated zinc oxide nanocatalysts. Nat. Commun. 2016, 7, 12273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Xie, J.; Jin, R.; Li, A.; Bi, Y.; Ruan, Q.; Deng, Y.; Zhang, Y.; Yao, S.; Sankar, G.; Ma, D.; et al. Highly selective oxidation of methane to methanol at ambient conditions by titanium dioxide-supported iron species. Nat. Catal. 2018, 1, 889–896. [Google Scholar] [CrossRef]
  20. Zhou, Y.; Zhang, L.; Wang, W. Direct functionalization of methane into ethanol over copper modified polymeric carbon nitride via photocatalysis. Nat. Commun. 2019, 10, 506. [Google Scholar] [CrossRef] [Green Version]
  21. Ji, P.; Zhang, J.; Chen, F.; Anpo, M. Study of Adsorption and Degradation of Acid Orange 7 on the Surface of CeO2 under Visible Light Irradiation. Appl. Catal. B 2009, 85, 148–154. [Google Scholar] [CrossRef]
  22. Feng, T.; Wang, X.; Feng, G. Synthesis of Novel CeO2 Microspheres with Enhanced Solar Light Photocatalyic Properties. Mater. Lett. 2013, 100, 36–39. [Google Scholar] [CrossRef]
  23. Chen, F.; Cao, Y.; Jia, D. Preparation and Photocatalytic Property of CeO2 Lamellar. Appl. Surf. Sci. 2011, 257, 9226–9231. [Google Scholar] [CrossRef]
  24. Sabari Arul, N.; Mangalaraj, D.; Kim, T.W.; Chen, P.C.; Ponpandian, N.; Meena, P.; Masuda, Y. Synthesis of CeO2 Nanorods with Improved Photocatalytic Activity: Comparison between Precipitation and Hydrothermal Process. J. Mater. Sci. Mater. Electron. 2013, 24, 1644–1650. [Google Scholar] [CrossRef]
  25. Chan, S.H.S.; Yeong Wu, T.; Juan, J.C.; Teh, C.Y. Recent Developments of Metal Oxide Semiconductors as Photocatalysts in Advanced Oxidation Processes (AOPs) for Treatment of Dye Waste-Water. J. Chem. Technol. Biotechnol. 2011, 86, 1130–1158. [Google Scholar] [CrossRef]
  26. Bamwenda, G.R.; Arakawa, H. Cerium Dioxide as a Photocatalyst for Water Decomposition to O2 in the Presence of Ce aq4+ and Fe aq3+ Species. J. Mol. Catal. A Chem. 2000, 161, 105–113. [Google Scholar] [CrossRef]
  27. Bamwenda, G.R.; Uesigi, T.; Abe, Y.; Sayama, K.; Arakawa, H. The Photocatalytic Oxidation of Water to O2 over Pure CeO2, WO3 and TiO2 Using Fe3+ and Ce4+ as Electron Acceptors. Appl. Catal. A 2001, 205, 117–128. [Google Scholar] [CrossRef]
  28. Primo, A.; Marino, T.; Corma, A.; Molinari, R.; García, H. Efficient Visible-Light Photocatalytic Water Splitting by Minute Amounts of Gold Supported on Nanoparticulate CeO2 Obtained by a Biopolymer Templating Method. J. Am. Chem. Soc. 2011, 133, 6930–6933. [Google Scholar] [CrossRef]
  29. Abanades, S.; Flamant, G. Thermochemical hydrogen production from a two-step solar-driven water-splitting cycle based on cerium oxides. Sol. Energy 2006, 80, 1611–1623. [Google Scholar] [CrossRef]
  30. Meiqing, S.; Xinquan, W.; Yuan, A.; Duan, W.; Minwei, Z.; Jun, W. Dynamic Oxygen Storage Capacity Measurements on Ceria-Based Material. J. Rare Earths 2007, 25, 48–52. [Google Scholar] [CrossRef]
  31. Melchionna, M.; Fornasiero, P. The role of ceria-based nanostructured materials in energy applications. Mater. Today 2014, 17, 349–357. [Google Scholar] [CrossRef]
  32. Karran Woan, Y.-Y.T.W.S. Synthesis and characterization of luminescent cerium oxide nanoparticles. Nanomedicine 2010, 5, 233–242. [Google Scholar]
  33. Muduli, S.K.; Wang, S.; Chen, S.; Ng, C.F.; Huan, C.H.A.; Sum, T.C.; Soo, H.S. Mesoporous cerium oxide nanospheres for the visible-light driven photocatalytic degradation of dyes. Beilstein J. Nanotechnol. 2014, 5, 517–523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Choudhury, B.; Chetri, P.; Choudhury, A. Oxygen defects and formation of Ce3+ affecting the photocatalytic performance of CeO2 nanoparticles. RSC Adv. 2014, 4, 4663–4671. [Google Scholar] [CrossRef]
  35. Gao, H.; Yang, H.; Yang, G.; Wang, S. Effects of oxygen vacancy and sintering temperature on the photoluminescence properties and photocatalytic activity of CeO2 nanoparticles with high uniformity. Mater. Technol. 2018, 33, 321–332. [Google Scholar] [CrossRef]
  36. Zhou, X.D.; Huebner, W. Size-induced lattice relaxation in CeO2 nanoparticles. Appl. Phys. Lett. 2001, 79, 3512–3514. [Google Scholar] [CrossRef]
  37. Choudhury, B.; Choudhury, A. Ce3+ and oxygen vacancy mediated tuning of structural and optical properties of CeO2 nanoparticles. Mater. Chem. Phys. 2012, 131, 666–671. [Google Scholar] [CrossRef]
  38. Choudhury, B.; Chetri, P.; Choudhury, A. Annealing temperature and oxygen-vacancy-dependent variation of lattice strain, band gap and luminescence properties of CeO2 nanoparticles. J. Exp. Nanosci. 2015, 10, 103–114. [Google Scholar] [CrossRef] [Green Version]
  39. Mogensen, M.; Sammes, N.M.; Tompsett, G.A. Physical, chemical and electrochemical properties of pure and doped ceria. Solid State Ion. 2000, 129, 63–94. [Google Scholar] [CrossRef]
  40. Goubin, F.; Rocquefelte, X.; Whangbo, M.H.; Montardi, Y.; Brec, R.; Jobic, S. Experimental and theoretical characterization of the optical properties of CeO2, SrCeO3, and Sr2CeO4 containing Ce4+ (f0) ions. Chem. Mater. 2004, 16, 662–669. [Google Scholar] [CrossRef]
  41. Shi, S.; Ke, X.; Ouyang, C.; Zhang, H.; Ding, H.; Tang, Y.; Tang, W. First-principles investigation of the bonding, optical and lattice dynamical properties of CeO2. J. Power Sources 2009, 194, 830–834. [Google Scholar] [CrossRef]
  42. Murali, A.; Lan, Y.P.; Sohn, H.Y. Effect of oxygen vacancies in non-stoichiometric ceria on its photocatalytic properties. Nano Struct. Nano Objects 2019, 18, 100257. [Google Scholar] [CrossRef]
Figure 1. The typical photographs of CeO2-raw and as-prepared CeO2-x catalysts at different calcination temperatures in Ar.
Figure 1. The typical photographs of CeO2-raw and as-prepared CeO2-x catalysts at different calcination temperatures in Ar.
Catalysts 10 00196 g001
Figure 2. (A) X-ray diffraction (XRD) patterns of CeO2-raw and CeO2-x; (B) comparison of particle size based on the calculation using the Scherrer equation on the basis of the (111) peak.
Figure 2. (A) X-ray diffraction (XRD) patterns of CeO2-raw and CeO2-x; (B) comparison of particle size based on the calculation using the Scherrer equation on the basis of the (111) peak.
Catalysts 10 00196 g002
Figure 3. (A) Ce 3d X-ray photoelectron spectroscopy (XPS) spectra, and (B) the relatively intensity ratios of CeIII/CeIV of CeO2-raw and CeO2-x.
Figure 3. (A) Ce 3d X-ray photoelectron spectroscopy (XPS) spectra, and (B) the relatively intensity ratios of CeIII/CeIV of CeO2-raw and CeO2-x.
Catalysts 10 00196 g003
Figure 4. UV–Vis diffuse reflectance spectra (DRS) of CeO2-raw and CeO2-x. Inset I: Tauc plots based on the DRS spectra, where [F(R)hν]0.5 is the Kubelka–Munk function for indirect-band semiconductors and is photon energy; Inset II: optical band-gap trend.
Figure 4. UV–Vis diffuse reflectance spectra (DRS) of CeO2-raw and CeO2-x. Inset I: Tauc plots based on the DRS spectra, where [F(R)hν]0.5 is the Kubelka–Munk function for indirect-band semiconductors and is photon energy; Inset II: optical band-gap trend.
Catalysts 10 00196 g004
Figure 5. Photocurrent curves of CeO2-raw and CeO2-x catalysts under the simulated solar light (a Xe lamp with an air mass (AM) 1.5 filter) irradiation with a 0-mV bias (vs. Ag/AgCl). The arrows show the light on and light off.
Figure 5. Photocurrent curves of CeO2-raw and CeO2-x catalysts under the simulated solar light (a Xe lamp with an air mass (AM) 1.5 filter) irradiation with a 0-mV bias (vs. Ag/AgCl). The arrows show the light on and light off.
Catalysts 10 00196 g005
Figure 6. Production rates of the photocatalytic products aldehyde (CH3CHO) and ethanol (CH3CH2OH) over CeO2-raw and CeO2-x photocatalysts under simulated sunlight irradiation. The ethanol selectivity is shown with orange dots.
Figure 6. Production rates of the photocatalytic products aldehyde (CH3CHO) and ethanol (CH3CH2OH) over CeO2-raw and CeO2-x photocatalysts under simulated sunlight irradiation. The ethanol selectivity is shown with orange dots.
Catalysts 10 00196 g006

Share and Cite

MDPI and ACS Style

Du, J.; Chen, W.; Wu, G.; Song, Y.; Dong, X.; Li, G.; Fang, J.; Wei, W.; Sun, Y. Evoked Methane Photocatalytic Conversion to C2 Oxygenates over Ceria with Oxygen Vacancy. Catalysts 2020, 10, 196. https://doi.org/10.3390/catal10020196

AMA Style

Du J, Chen W, Wu G, Song Y, Dong X, Li G, Fang J, Wei W, Sun Y. Evoked Methane Photocatalytic Conversion to C2 Oxygenates over Ceria with Oxygen Vacancy. Catalysts. 2020; 10(2):196. https://doi.org/10.3390/catal10020196

Chicago/Turabian Style

Du, Jin, Wei Chen, Gangfeng Wu, Yanfang Song, Xiao Dong, Guihua Li, Jianhui Fang, Wei Wei, and Yuhan Sun. 2020. "Evoked Methane Photocatalytic Conversion to C2 Oxygenates over Ceria with Oxygen Vacancy" Catalysts 10, no. 2: 196. https://doi.org/10.3390/catal10020196

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